Sei sulla pagina 1di 147

The Morphology of Coating-Substrate Interface

in Hot-Dip-Aluminized Steels

By
GUL HAMEED AWAN
2001-PhD-Met-01

Department of Metallurgical & Materials Engineering


University of Engineering and Technology
Lahore-Pakistan
The Morphology of Coating-Substrate Interface in Hot-Dip-
Aluminized Steels

A thesis submitted to the University of Engineering and Technology, Lahore as a


partial requirement for the degree of DOCTOR OF PHILOSPHY in Metallurgical &
Materials Engineering

Approved on ________________________________

Internal Examiner _________________________________


Dr Faiz ul Hasan
Professor
Department of Metallurgical and Materials Engineering
UET, Lahore.

External Examiner ___________________________________


Professor Dr.M. Saleem Shuja
Rector
University of Lahore

__________________________________
Chairman

________________________________
Dean Faculty of CMME
This thesis was evaluated by the following Examiners:

External Examiners:

From Abroad:
Dr John Pilling
Adjunct Associate Professor,
Materials Science and Engineering
Michigan Technological University, USA

Dr Gordon W Lorimer
Emiritus Professor of Physical Metallurgy and Materials Science,
School of Materials,
The University of Manchester, U.K.

From Pakistan:
Prof Dr Muhammad Saleem Shuja
Rector, University of Lahore.

Internal Examiner:
Dr Faiz ul Hasan
Professor, Department of Metallurgical and Materials Engineering,
University of Engineering and Technology, Lahore.
DEDICATION

This thesis is dedicated to my parents, wife Hummaira, daughter


Ainee and sons Muaaz, Umer & Abdullah.
ACKNOWLEDGEMENT

All and every kind of praise is for Allah Who guides us from darkness to light and helps
us out of our difficulties.

I am highly indebted to the Vice Chancellor UET Gen. Akram Khan and Dr. K. E.
Durrani (Director Research) for their continuous encouragement during the period of my
PhD research.

I am extremely grateful to my research supervisor Prof Dr Faizul Hasan for his keen
interest in my research work, expert guidance, valuable discussions and encouragement.

I would like to express my sincerest gratitude to Mr. Furqan Ahmed (Lecturer


Department of Metallurgical and Materials Engineering UET Lahore) for his continuous
help and invaluable advices during my research work.

Prof Dr Khaleeq Ur Rehman Chairman Physics Department and Prof Dr Khalid


Mahmood Ghouri have kindly extended continuous excess to the practical facilities
available with them in their laboratories.

I extend my admiration and appreciation to Prof Dr Liaqat A Kasuri to share with me his
valuable academic information during PhD research.

I am thankful to the Laboratory staff at Department of Metallurgical and Materials


Engineering for their valuable cooperation and support specially Mr. Abdul Qayyum, Mr.
Khalid and Mr. Inderias.

The technical assistance provided by Mr. Siddique and Mr. Atique in making the
specimens and the bending machine is hereby acknowledged.

Gul Hameed Awan


Table of Contents

Abstract 1

1. Literature Survey 2
1.1 What is Aluminizing 2
1.2 Aluminizing methods 2
1.3 Hot Dip Aluminizing 4
1.4 Fe-Al Phase Diagram 14
1.5 Mechanism of Hot dip Aluminizing 20
1.6 Role of Intermetallics 21
1.7 Studies of the interlayer 22
1.8 Control of intermetallics at interface 25

2. Experimental 43
2.1 Size Shape and Chemical Composition of Steel Specimens 43
2.2 Preparation of Aluminum Melts 44
2.3 Cleaning / Fluxing prior to Aluminizing 45
2.4 Bending of Aluminized Specimens 48
2.5 Metallographic Examination of Aluminized Specimens 48
2.6 X-ray Diffraction of Aluminized Samples 49
2.7 Scanning Electron Microscopic Examination 49
2.8 Tensile Testing of Al-Si alloy samples. 49

3. Results and Discussion 50


3.1 Aluminizing in Pure Aluminum 50
3.2 X-ray Diffraction of aluminized samples 61
3.3 Obtaining Powdered samples of the interlayer 70
3.4 Aluminizing in Al-Si alloys 71
3.5 XRD and SEM of samples aluminized in Al-Si melts 74
3.6 Lattice Parameter Measurements of Fe3Al and Fe2Al5 88
3.7 Bending of Aluminized samples 89
3.8 Effects of ‘over-modification’ of Aluminizing Melts 103

4. Conclusions 112

5. Suggestion for Future Work 113

6. Appendix-I 114
Appendix-II 117

7. References 120
ABSTRACT

In Hot-Dip-Aluminized steels, the morphology and the profile of the interface between
the aluminum coating and the substrate steel, are affected both by the composition of the
molten aluminum as well as by the composition, and even the microstructure, of the
substrate steel. This effect has been investigated using optical and scanning electron
microscopy, and X-ray diffraction. The reaction between the steel and the molten
aluminum leads to the formation of Fe-Al intermetallic compounds on the steel surface.
The thickness and the morphology of the interlayer vary with the silicon-content of the
molten aluminum. In hot-dip-aluminizing with pure aluminum, the interlayer is ‘thick’
and exhibits a finger-like growth into the steel. With a gradual addition of silicon into the
aluminum melt, the thickness of the interlayer decreases and the interface between the
interlayer and the substrate successively becomes ‘smoother’. With an increase in the
carbon-content of the substrate steel the growth of the interlayer into the steel is impeded
by the pearlite phase, whereas the ferrite phase appears to dissolve more readily. X-ray
diffraction and scanning electron microscopic studies showed that the interlayer formed
in samples aluminized in pure aluminum, essentially consisted of orthorhombic Fe2Al5,
while with a gradual addition of silicon into the aluminum melt, a cubic phase based on
Fe3Al also started to form in the interlayer and replaced most of the Fe2Al5. It was further
observed that the grains of Fe2Al5 phase exhibited a preferred lattice orientation, and also
that even when other phases are present in the interlayer, the phase at the transformation
front is always the Fe2Al5.

Bending experiments carried out on aluminized specimens showed that as the thickness
of the inter-metallic layer increased the angle, at which the start of the cracking in the
aluminum coating during bending was observed, decreased. Metallographic examination
of bent samples showed that the interlayer developed cracking much earlier than the
appearance of the cracks on the surface. These experiments suggested that the aluminized
steel flats (or sheets) exhibit very limited formability.
1 LITERATURE SURVEY

1.1 What is Aluminizing?

Aluminizing is a process in which the surface of a metallic component is coated with a


layer of aluminum. Steels and nickel-based alloys are the most common metals that are
aluminized for commercial applications. Aluminum coating on steel or other materials
increases their resistance to oxidation at high temperatures (500-800°C). It may also
increase the corrosion-resistance of steel in hydrocarbons and sulphurous atmospheres.

1.2 Aluminizing Methods

Various methods can be employed to obtain an aluminum coat on the surface of metals.
In some cases the surface may only be enriched with aluminum by impregnation to
enhance its resistance to oxidation at elevated temperatures. The various methods
commercially employed for aluminizing are as follows [1].

1) Pack Aluminizing
2) Spray Aluminizing (Metallizing)
3) Vacuum Aluminizing
4) Gas Aluminizing
5) Cladding
6) Electrolytic Coating
7) Hot Dip Aluminizing

Pack Aluminizing [1] is carried out at fairly high temperatures (900-1080°C). The
surface to be aluminized is first cleaned of the scale and dirt and then packed in air-tight
retorts with ‘aluminizing mixture’, which consists of aluminum or ferro-aluminum dust
or powder. The retorts are then heated to the aluminizing temperature for upto 30 hrs.
This method is expensive and time-consuming and is only recommended for articles of
intricate shapes. Typical applications of pack-aluminizing are nickel-based alloys used in
gas turbines, although steels can also be pack-aluminized to improve its high temperature
oxidation resistance.

Spray Aluminizing (Metallizing). In this method an aluminum coating of a precise


thickness is sprayed on pre-cleaned surface [1]. The process includes melting and
spraying the metal with a jet of compressed air (at 25-60 psi pressure) with special
metallizers. The bond thus obtained is of low strength, although the adherence can be
improved by roughing the surface of the substrate. Annealing at 850-1250°C for long
time is done to increase the bond strength.

Vacuum Aluminizing. In this process an aluminizing coating is applied by vaporizing


the aluminum and condensing it on the substrate. The coating thickness in this case is of
the order of 0.1 micron [1]. Vacuum chamber with a pressure of 10-3 - 10-5 mm of Hg is
required. Electron beam devices are used to melt the aluminum and raise its temperature
up to 1400°C. Substrate also has to be heated up to 175-370°C. This process gives a good
coating adherence.

Gas Aluminizing. In the gas-aluminizing method, the steel substrate is impregnated with
aluminum by means of gaseous phase of aluminum chloride [1]. The process is carried
out in a retort which contains the mixture of following composition
45% pure aluminum
45% Al2O3
10% AlCl3
This mixture in the retort is heated to 600°C. The substrate to be aluminized is placed at
the opposite end of the retort and is heated there at 900-1000°C. Impregnation of steel
takes place according to the following reaction:
AlCl3 + Fe  FeCl3 + Al
Due to the complexity of the process and a considerable consumption of energy, this
method is rarely used in industry.

Cladding is accomplished by cold-rolling together the sheets of steel and aluminum.


During the rolling process metallic bond is formed between aluminum and steel. In this
way bimetallic strip or sheets are also formed [1].

Electrolytic Coating. In electrolytic coatings either fused electrolytes composed of a


mixture of fused salts of aluminum chloride or aluminum in ethyl bromide and benzene
are used. Before coating the substrate is thoroughly cleaned, degreased and pickled in
HCl solution [1]. The deposition-rate of aluminum metal onto the substrate in this process
is very slow; typically a coating thickness of 0.01 mm may be obtained in around 30
minutes.

1.3 Hot Dip Aluminizing

In this method of aluminizing the steel substrate is dipped after cleaning the surface into
molten aluminum or an alloy of aluminum at 700-800°C for a certain length of time. The
process is essentially a thermo-chemical treatment similar to ‘galvanizing’ or zinc-
coating. During the holding of substrate in molten aluminum a reaction takes place
between solid iron and liquid aluminum producing one or more inter-metallic compounds
of FenAlm type between the outer aluminum layer and the substrate steel [1]. The
thickness of the interlayer depends on holding time and temperature. The quality of
aluminizing is markedly affected by the cleanliness of the surface of substrate. The
process was developed after 1st world war in Russia, U.S.A. and Japan.

1.3.1 Continuous HDA of Mill Products

Aluminum hot dip coating of steel strip and wire is performed by a number of
patented processes, using continuous anneal-in-line equipment similar to that used for
galvanizing. The process consists essentially of three operations:

 surface preparation,
 pre-heating of the steel base,
 aluminum coating.
Surface preparation is a two-phase operation. First, all soil is removed from the surface
by oxidizing at elevated temperature or by chemical cleaning. Then, the surface oxides
are reduced in a suitable atmosphere to prepare the strip for coating [1].

Because the reaction between aluminum and steel is extremely rapid, the immersion time,
the temperature of the molten aluminum, and the temperature of the strip before and after
coating must all be controlled to prevent the formation of an excess of iron aluminum
interfacial alloy. Unless a void layer separates the alloyed coating from the base metal,
the amount of iron in the alloyed coating increases with time as the aluminum continues
to diffuse into the base metal.

The amount of brittle interfacial alloy layer can be controlled also by the addition of
silicon to the coating bath [1]. This increases the apparent ductility of the coating,
enabling more severe fabrication of the sheet without peeling of the coating. The
thickness of the interfacial layer decreases rapidly as the silicon content increases to
about 2.5%. A smaller decrease occurs as the silicon content is further increased.

1.3.1.1 Equipment

Equipment for a continuous line of hot dip aluminum coating consists of a feeding
section, a furnace section, and a delivery section. In the feeding section, equipment
uncoils incoming strip and feeds it into the coating line at a designated constant speed
under specified tension. The furnace section contains the preheating or oxidizing furnace,
the annealing furnace, the cooling furnace, and the coating pot. If chemical cleaning is
used, alkaline cleaner and water rinse tanks are substituted for the preheating furnace [1].
The cooling furnace is connected directly with the annealing furnace and extends to the
coating bath with its end sealed by means of a snout extending into the molten aluminum
bath. A dry reducing atmosphere of hydrogen and nitrogen is maintained within the
annealing and cooling furnaces. The delivery section is equipped to provide rapid cooling
and sufficient time for setting the coating before the strip contacts the support roll over
the coating bath. Drive rolls and equipment for looping roller leveling, coiling and
shearing, and stretch leveling and surface conditioning are all contained in the delivery
section. The Process line is shown in Figure 1.1.

Figure 1.1. Process diagram of continuous hot dip aluminizing of steel strip and wire [2].

1.3.1.2 Effect of Coating on Strength and Fabricability

For different reasons, aluminum-coated steel strip and wire decreases measurably in
strength. The strength of strip decreases because it is normally annealed prior to coating
in order to improve its fabricability. Wire strength decreases as a result of the high
temperature, above 650°C, of the hot dip coating bath [1]. An example is aluminum
conductor steel reinforced ACSR wire, which is either aluminum coated or galvanized.
Its tensile strengths range from 1140 to 1450 MPa (165 to 210 ksi) depending on the type
and class of coating specified.

Steel sheet coated with aluminum silicon alloy withstands moderate forming, drawing
and spinning operations without flaking or peeling of the coating. Steel sheet coated with
commercially pure aluminum withstands moderate brake and roll forming operations and
can be spun or embossed, but is not suitable for drawing [1]. Sheet with either type of
coating can be given a 180° bend around a diameter equal to twice the thickness of the
material; however, in any forming operation, it is advisable to allow liberal radii to
prevent crazing of the coating.
Because sheared edges are susceptible to corrosion, the use of aluminum coated sheet for
fabricated assemblies may be limited for appearance considerations. Corrosion protection
is not impaired, because there is no undercutting of the coating.

Coated sheet should be fabricated before the coating is diffused. Diffusion converts the
coating to an iron-aluminum compound, which is very brittle [1]. The alloy arises from
interaction of molten, commercially pure aluminum with the steel surface and is the agent
that bonds the total coating to the substrate [2]. The alloy is an essential part of the
coating protection system, supplementing the aluminum layer and providing a second line
of defense to ensure long term durability. Control of the alloy layer thickness and
thickness uniformity assures the degree of coating formability.

Aluminum spontaneously forms an aluminum oxide passive film, as it is a characteristic


of aluminum. This film imparts its usual resistance to major environmental factors
influencing corrosion behavior in water and soils [2]. Corrosion due to dissolved gases
and carbon dioxide and erosion corrosion due to high velocity water are the common
influential factors in a pipe waterside environment. The passive film imparts high
resistance to all of these factors. Figure 1.2 shows the typical microstructure of a HDA
steel sheet.
Figure 1.2. Microstructure of a hot-dip-aluminized steel sheet showing; surface
Al-coating, Al-alloy interlayer, and the steel substrate [2].

1.3.1.3 Typical Applications

A wide variety of industrial, farm, and consumer products are fabricated from steel sheet
aluminum coated at the mill. The following products require resistance to oxidation and
corrosion at temperatures from 95 to 680°C: combustion chamber and outer casings,
agricultural crop dryers, automotive mufflers, space heaters, furnace flues, oven interiors,
barbecue grills, and wrappers for water heater elements. Fabrication of these parts
requires moderate drawing, forming, punching and spot welding [1].

1.3.2 Batch Hot Dip Coating of Parts

Steel parts to be aluminized are first cleaned by hot-alkaline dipping and water rinsing.
Steel parts are then de-scaled by abrasive blasting or acid pickling, rinsed and dried [1].
Gray or malleable iron parts are given an additional cleaning in molten salt to remove
carbon smut.

For coating parts by the fused-salt fluxing method, an electrically heated fluxing furnace
is used in conjunction with a coating furnace. The fluxing furnace is lined with a porous
refractory brick, such as mullite, and insulated with porous silica brick surrounded by a
steel shell. The brick lining must be compatible with both aluminum and halide salts of
sodium and potassium, at temperatures up to 790°C. The fluxing furnace contains alumi-
num for maintenance of the salt bath.

The coating furnace is preferably a low-frequency induction furnace with a mono-cast or


rammed lining. To prevent oxidation, a layer of molten salt 25 to 75 mm thick floats
above the molten aluminum [1]. A cast iron pot can be used for the coating furnace, but
the pot must be coated with a wash of iron oxide and titanium dioxide in a silicate binder
to prevent attack by the molten aluminum.

1.3.2.1 Procedures and Control

An example of an automatic conveyorized line for high-production batch hot dip coating
of small parts such as fasteners is shown in Figure 1.3. The procedure consists of
cleaning, preheating, fluxing, and coating [1].

Figure 1.3. Flow sheet diagram showing batch hot-dip-aluminizing process for steel
parts [1].

In this cleaning process, about 100 kg load of parts is put into a basket and immersed in
reducing salt at a temperature of 540°C for 20 to 25 min. The parts are dip rinsed in cold
water and then pickled for 15-20 minutes in 8-10% H2S04 at 70°C. The parts are dipped
and sprayed with cold water, and dried in circulating hot air.

For the preheating and fluxing processes, a basket is loaded with 10-20 kg (20 to 40 lb) of
cleaned parts. The parts are dipped in salt (40% NaCl, 40% KCI, 10% AlF3, and 10%
Na3AlF6) for 8 to 15 min at 705°C [1]. The basket is in constant motion during the
preheating, fluxing and coating processes to remove any trapped air and to flux all
surfaces of the part thoroughly.

The basket of parts proceeds to the coating process where the parts move through the
molten aluminum alloy at 700°C for 1 to 2.5 min. Excess aluminum is removed from
parts by centrifuging, air blasting or shaking. After the basket is unloaded, the parts are
quenched in water at 70 to 80°C. Then, they are air dried and inspected.

In another coating process, the pickling and fluxing operations are eliminated. Parts are
abrasive grit blasted and then immersed directly into molten aluminum. The bath surface
is skimmed before parts are immersed and before they are removed. Unless it is carefully
controlled, this procedure has the disadvantages of (a) heating parts to coating
temperatures in the molten aluminum, producing a heavier alloy layer, and (b) dissolving
additional iron into the bath.

An iron content in excess of 2%; may produce rough coatings. The iron content of the
coating bath can be controlled to some extent by reducing bath temperature. At lower
temperatures, aluminum high in iron becomes mushy and concentrates along the sides
and corners of the pot from which it can be removed rapidly [1]. Another method of
reducing the iron content is to remove part of the bath and add iron-free aluminum.

Temperature control of the molten aluminum is important. Higher temperature increases


the thickness of the intermetallic layer and decreases the thickness of the pure aluminum
overlay. A wider range of temperatures can be tolerated when the aluminum contains 5%
silicon.
Immersion time must he controlled closely. The thickness of the intermetallic layer
increases with immersion time. The pure aluminum overlay, is not affected, because its
thickness depends upon the viscosity of the aluminum and on the speed with which the
part is withdrawn from the bath [1].

1.3.2.2 Limitations

Small parts with fine threads are considered impractical to coat by dip methods. Washers
and cup shaped parts that may nest offer some difficulty in batch processing. Continuous
uniform coatings are difficult to produce on items with complicated configurations
involving blind holes and re-entrant angles, in which air can be entrapped [1].

Long slender parts of thin cross section may warp at coating temperatures 660 to 720°C.
Such parts must be well supported or immersed in the vertical position to minimize
warping; however, some deformation still may occur as a result of the stress-relieving
effect of the bath temperature. The use of the lower temperature aluminum-silicon alloy
may help to minimize distortion.

The strength of coated parts, especially those made of cold worked material, may be
reduced by the coating operation. In certain instances, this condition can be alleviated by
the use of higher strength material. One producer of aluminum-coated bolts made of 4140
steel states that a tensile strength of 760 to 790 MPa is the maximum practical strength
that can be obtained after coating without resorting to high-alloy steels. The composition
of the steel does not limit the batch hot dip coating process. The choice of steel depends
on the strength and service requirements.

Aluminum coatings applied by hot dipping are more costly than hot dip coatings of other
metals. This is because the basis metal- requires more thorough cleaning, more heat is
needed for coating at about 700°C, and electricity must be used for heating the ceramic
pot [1]. Fuel costs may be decreased by the use of iron pots fired by gas or oil, but
ceramic pots are preferred for this process.
1.3.3 Coating of High Production Parts

Automotive poppet valves, blades and nozzle vanes for gas turbine engines, and fasteners used
in connection with pole line hardware and aluminum assemblies are examples of high
production parts that are coated with aluminum [1].

1.3.3.1 Corner Castings

In contrast to the complexity of the high-production line for fasteners and complicated
small parts is the procedure for coating large corner castings used on aluminum
fabricated cargo containers. These castings, each weighing about 14 kg (30 lb), are made
of medium-carbon steel. The procedure for aluminum coating a production run of 12,000
of these castings is as follows:

 Inspection: incoming castings are inspected visually for flaws. If cavities are
present, a fixture is used to check their size. Because the finished product must mate
with another steel fixture, the size of part and buildup of excess aluminum are
critical.
 Cleaning: castings are oven-baked in batches at 200°C for 5 h to remove gases and
grease contamination, and then blasted with clean steel grit, 25 to 40 mesh.
 Racking: castings are handled with clean cotton gloves to prevent surface
contamination during racking. Castings must be racked so that air pockets do not
develop during immersion in molten aluminum. An air pocket causes oxidation of
the ferrous metal surface and interferes with coating.
 Immersion: each rack containing 30 pieces is dipped into molten aluminum, 99.5%,
at 720°C for 12 min. The iron and silicon contents of the bath are determined
weekly. When the sum of the two exceeds 3% uncontaminated aluminum is added
to reduce the concentration, or the bath is scrapped and recharged.
 Cleaning: immediately on removal from the molten aluminum, the rack is shaken
and the castings are air blasted to remove excess aluminum.
 Inspection: each piece is inspected to ensure fit [1].
1.3.3.2 Turbine Blades and Vanes

These parts, usually made of super-alloys for use at 700 to 1100°C, can be coated
satisfactorily or impregnated with aluminum by the slurry method, by pack diffusion, or
by hot dipping and diffusion treatments [1]. The earliest work used a hot dip process that
recognized the importance of a thin alloy surface layer for optimum resistance to thermal
shock and corrosion.
The following are basic steps in the processing of turbine blades or turbine vane segments
by the hot-dip-and-diffuse method.

 Surface preparation:

o Immerse in molten caustic, oxidizing at 500°C for 15 min


o Rinse in water at 80°C for 2to 3 min
o Wet blast with 240-mesh grit at gage pressure of 690 kPa

 Aluminum coating:

o Immerse in molten salt flux: 35 to 37% KCl, 35 to 45% NaCI and 0.5 to 12%
AlF, 8 to 20%, Na3AlF6 at 720°C for 3 to 5 min. This flux should be maintained
above the molten aluminum for the purpose of activation.
o Immerse in aluminum bath at 700oC for 3 to 10 s.
o Rinse in molten flux, then air blast or centrifuge and wash in water.
o Dip in 25% HNO3 solution at 21oC to brighten and clean the surface.
o Leach in 10 to 12% HCl solution at 65 to 75°C when necessary to remove
aluminum.
o Wash in water at 100°C and dry

 Heat treating for diffusion:

o Load in furnace at temperature


o Diffuse at 1090 to 1150°C for 2 h at temperature in air, argon, helium or
endothermic generator gas.
o Cool slowly in furnace to below 760ºC

Final cleaning of the diffusion heat treated blades is necessary if an accurate dye
penetrant or fluorescent oil penetrant inspection is required [1]. Wet blasting is suitable
for removing the oxide prior to such inspection.

1.3.3.3 Automotive Poppet Valves

The most advanced method of coating poppet valves consists of these operations:

 Clean and degrease after finish machining and grinding

 Induction preheat surfaces to be coated

 Spray aluminum onto preheated surfaces. Spray gun should be at a fixed position
from the preheating station. Valves are rotated to obtain uniform mechanically
bonded coating.

 The rotating valve progresses to an induction heating station, where the coating is
heated to bond it metallurgically to the substrate metal by diffusion to a depth of
about 25 μm (1 mil).

 Cool in air cooling chamber


 Inspect visually and metallurgically

This method results in minor discoloration of the valve head and a slight roughening of
the aluminum coating. Subsequent surface conditioning is unnecessary, and even
undesirable, because the full thickness of heat resistant surface alloy is an advantage
during service. The slightly roughened surface, consisting of undiffused aluminum,
becomes smooth in the first few seconds of engine operation [1]. Most aluminum-coated
valves are made by this method, which is the most economical and the most effective for
developing soundness, uniformity and durability.
Valves may be coated also by the hot-dip-and-diffuse method or by spraying and
subsequently diffusing in a salt bath, but the cost of either of these procedures is greater
than for the method described above.

1.4 Fe-Al Binary Phase Diagram

1.4.1 The Range 0-54 At. % Al

The liquid and solidus curve of the α phase in the Figure 1.4 was plotted using the highest
liquidus temperature and the smallest solidification intervals reported [4-7].

The vertex of the γ loop has been reported as about 1.2 wt % (2.4 at. %) Al [6], and about
1.0 wt % (2.0 at. % ) Al [8]; see data points in the insets of Figure 1.4. The γ/(α + γ) and
α/(α + γ) boundaries calculated by [9] extended to about 1.2 at. % (0.6 wt %) and about
2.0 at. % (1.0 wt %) Al, respectively, at about 1150°C.

Because of the lack of reliable experimental data, the location of the high-Fe boundary of
the α + ε and α + δ field is uncertain. In addition, information related to the probable
connection of the latter phase boundary with the disorder → order transformation α → β 2
= FeAl is still missing. The α/(α + ε) boundary is merely based on the thermal results of
Gwyer and Phillips [7]; its extension to lower temperature is tentative. Both [6] and [10]
found that solubility of Al in α-Fe decreased with fall in temperature. It is not known
whether the composition FeAl lies inside or outside the heterogeneous field.
Figure 1.4. Iron-Aluminum binary phase diagram [3].
Table 1.1. Description of phases formed at different wt% Al [3].

1.4.1.1 The Order-Disorder and Magnetic Transformation.

First indication of the occurrence of transformation in the α solid solution was the
detection of an ordered b.c.c. structure of the CsCl (B2) type [11]. The existence of two
super lattices based on the ideal composition Fe3Al = β1 and FeAl = β2 was established by
Bradley and Jay [12,13]. Alloys with 0-51 at. % Al were (a) quenched from 700 and
600°C and (b) slowly cooled from 750°C. Upto about 18.5 at. % Al the atomic
distribution in the b.c.c. lattice of α-Fe was random, independent upon heat-treatment. In
the range 18.5-25 at. % Al the atomic arrangement was random after quenching from
600°C or higher, but ordered after slow-cooling to room temperature, with the degree of
ordering increasing to 92% at 25 at. % Al. The super-lattice Fe3Al gradually emerging
from the random arrangement is of the BiF3 (DO3) type, and the lattice constant twice
that of the α phase. Between 25 and 34 at. % Al the structure was of the Fe3Al type
(gradually changing to FeAl type) after slow cooling and of the FeAl (B2) type after
quenching from 600°C or higher temperature. Above 34 at. % the structure was solely of
the FeAl type, after both slow cooling and quenching from 600 and 700°C. These finding
indicate that the Fe3Al superlattice is formed below 600°C whereas the FeAl superlattice
is stable upto 700°C; however, the formation of the latter superlattice might not be
suppressed by quenching. With the heat treatments used in these studies, no regions
containing both α and Fe3Al structures and both Fe3Al and FeAl structures could be
found. Bradley and Jay [12,13] therefore claimed that the transition of one type of
structure to the other was continuous. Additional studies using clearly defined long-time
anneals, is necessary to prove whether this is true.

Some workers [14-17] studied the order-disorder transformations in the Fe3Al range by
means of thermal [15], specific-heat [15,17], dilatometric [15], resistivity [14,16,17], and
magnetic [15,17] methods. The transition temperatures reported by these workers are
shown in Figure 1.5, which also contains the curves of the curie points of the disordered
α-phase and the ordered Fe3Al structure based on the fairly closely agreeing results of
[15,16,18,19]. Curie temperatures were also determined by [6,20,17].

As shown in Figure 1.5, Saito [17] found the transition temperatures in the range 10-24
atomic % Al which were claimed to belong to ‘short-range ordered superstructure of
Fe13Al3 (18.75 atomic % Al)’. This type of structure was found to coexist with the Fe3Al
superstructure in alloys containing more than 21.4 at. % Al. A new phase diagram was
proposed for the range 10-30 atomic % Al.

Although considerable information on order-disorder and magnetic transformations is


available, more work is needed. As a consequence, the transformation curves and dashed
structure boundaries in Figure 1.4, proposed in principle by Bradley and Taylor [21], are
to be regarded as tentative. Transition temperatures of the order-disorder transformation
FeAl (β2) → α have not been determined as yet. For more complete information of the
transformation in the range 0-50 at. % Al, the papers [12-27] should be consulted.
Figure 1.5 Fe-rich-end of the Fe-Al phase diagram showing different ordering
and disordering transformations [28].

1.4.2 The Range 54-77 At. % Al

The solidification equilibria suggested by [4-7] are greatly at variance. The various
findings have been summarized by Hansen [28]. In contrast to [5, 6], who found only one
phase of variable composition between 68 and 76 at. % Al, some workers [7, 29, 30] gave
evidence of the existence of two phases in this region, Fe2Al5 = ε and FeAl3 = ζ. In
addition to the ε phase which forms peritectically at 1230°C [5], 1207°C [6], 1232°C [7]
or 1210°C [29] and decomposes eutectoidally at about 1100 [5], 1080 [6,29] or 1103°C,
another phase between 65 and 70 at. % Al was detected by Gwyer and Phillips [7]. This
phase, δ, was shown to be formed by the peritectoid reaction ε + ε → δ at 1158oC i.e.,
only about 7oC below the ε + ε eutectic [7]. Because of the incompleteness of this
reaction on relatively rapid cooling and insufficient annealing, no single-phase alloy
could be obtained in this range. Gwyer and Phillips [7,31] assumed this phase to contain
about 65 at. % Al; however, Hansen [28] suggested it to be somewhat richer in Al and
correspond to the composition FeAl2.

The phase relationships shown in Figure 1.4 are based mainly on the work of Gwyer and
Phillips [7]. However instead of a liquidus maximum (1161°C) at FeAl3, and a eutectic
FeAl2-FeAl5 at 1159°C and about 74 at. % (54 wt %) Al, Ageew and Vher [10] suggested
the peritectic reaction: melt + Fe2Al5. The thermal data of [5] and the microstructure of
the alloys with 72.8 at. % Al [10] and 72.7 at. % Al [7], showing peritectic rather than
eutectic structures, are in favor of this generally accepted version.

The intermediate phase richest in Al has been widely accepted to correspond to the
composition FeAl3 [4,6,7,29]. Bradley and Taylor [30] claimed to have found strong X-
ray evidence that FeAl3 = ζ decomposed at some temperature below 600°C to give a
mixture of Fe2Al5 and Fe2Al7 (77.8 at. % Al), the latter being stable between about 77.5
and 78.6 at. % (62.5 and 64 wt %) Al. The powder pattern of Fe2Al7 was reported to
differ only slightly from that of FeAl3, the difference being clearly defined as second-
order effects [30]. According to Bradley and Taylor [21] no indication of thermal effect
was found on cooling curves, however, these results could not be reproduced by recent
careful X-ray studies [32]. There were no differences in the powder patterns of specimens
(including a single crystal) containing 76.4-76.8 at. % Al, quenched from 900°C or
subjected to a long-time anneal at 470°C. The conclusion that FeAl3 does not decompose
was also drawn by [33]. On the other hand, there is some indication that the composition
of the intermediate phase coexisting with the aluminum rich solid solution is lower in
iron than previously assumed: [34], who analyzed primary crystals extracted from alloys,
found the compositions to be close to Fe2Al7. Additional work is needed to verify these
and other findings [35,36] which indicate that the phase ‘FeAl3’ is actually richer in
aluminum, or to conclude that it has a wider homogeneity range. Until then, the formula
FeAl3 should be retained for convenient reference.
The boundaries of Fe2Al5 (ε phase) given by various workers [7, 29, 30, 37] differ
somewhat ranging from 53-55.5 wt % Al on the Fe-side to 55.5-57 wt % Al on the Al-
side. As a compromise, 53-56 wt % (70-72.5 at. %) Al was accepted in Figure 1.4.

1.4.3 The Range 77-100 At. % Al

The liquidus curve is based on the closely agreeing thermal data of [5,7,38]. The eutectic
temperature was given as 646-649°C [4], 648°C [39,7], 652°C [5], 650°C [40], 655°C
[41], 653°C [7] and 654oC [39,42] and the eutectic composition in wt % Fe as 2.0 [39]
2.5 [40] about 1.7 [41] about 2.2 [43] 1.9 [7], and 2.0 [38]. There appears to be a little
doubt that these differences are mainly caused by the use of aluminum of different
purities. Also, the tendency of the melt to under cool might have affected the results. The
most reliable data indicate the eutectic point to be located at 1.8 ± 0.1 wt % (0.9 ±0.05 at.
%) iron and 654-655°C.

Results as to the solid solubility of iron in aluminum were chiefly reported by [41,44,45].
Those of Edgar [45], based on lattice-spacing measurements, are most reliable and
complete; 0.052 wt % Fe at 655°C, 0.025% at. 600°C, 0.006% at. 500°C, and practically
zero at 400-450°C. With extremely high cooling rates during solidification,
supersaturated solutions containing as much as 0.17 wt. % Fe can be obtained [46].
1.4.3.1 Crystal structures

Lattice parameter of alloy in the range 0-50 at. % Al were reported by [47,29,13] in
which those of Bradley and Jay [13] being most reliable. The structural changes
connected with the disorder → order transformation α→Fe3Al (= β1) and α → FeAl
(= β2) were elucidated by [12,13]. The Fe3Al structure is of the BiF3 (DO3) type, with a
lattice constant twice that of the α phase, and the FeAl structure is isotypic with the CsCl
(B2) structure.
The ε phase was suggested to be b.c.c. with 16 atoms per unit cell and δ phase (FeAl2) to
be rhombohedral with 18 atoms per unit cell [29,48]. The structure of the later phase was
claimed to be more complicated [30]. The ε phase Fe2Al5 first believed to be monoclinic
with 56 atoms per unit cell [28], was reported to be orthorhombic, with a=7.675Å, b =
6.403Å, c=4.203Å for 72.0 ± 5 at. % Al [49].

The ζ phase (FeAl3) was previously believed to be orthorhombic [29,48,50,51]. A


complete structure analysis, using specimens with about 76.5 at. % Al revealed the
structure to be monoclinic with 100 atoms per unit cell and a = 15.489Å, b = 8.083Å,
c = 12.476Å, β = 107o43΄ ± 1΄ [52]. The existence of an unstable phase Fe2Al claimed to
be present in a complex Fe-Ni-Cr Alloy was reported by [53]. It was stated to be isotypic
with MgZn2.

1.5 Mechanism of Hot-Dip-Aluminizing

Hot-dip-aluminizing (HDA) involves essentially a reaction between solid Fe at steel


surface and liquid Aluminum. The steel substrate to be aluminized is immersed in a bath
of molten aluminum or its alloy. The coating process occurs by wetting, dissolution,
inter-diffusion, and then the formation of intermetallics. It must be appreciated that in hot
dip aluminizing it is difficult to join the aluminum with steel without the formation of
intermetallic layer at the face of the solid because of their limited mutual solubility. A
clean surface of the substrate ensures good wetting and authenticates better aluminizing.
Aluminum has good affinity with oxygen. The presence of iron oxide in the melt further
becomes the problem for hot-dip-aluminizing process.
The intermetallic layer is brittle in nature. This layer grows and dissolves concurrently
into the molten aluminum alloy, which is directly associated with the loss of the steel
substrate. The growth and the dissolution rates of the intermetallic layer determine
thickness of the layer. The rates are closely related with temperature of the molten
aluminum alloy and dipping time of the steel [54,55,56-58]. The thickness of the layer
also varies depending on chemical composition of the molten alloy [59] and of the steel
substrate.
1.6 Role of Intermetallics

Hot-dip-aluminized steels have got very good resistance against oxidation and corrosion
at room and elevated temperature and it depends upon the physical and mechanical
properties of the alloy, chemical bond between the aluminum and steel [60].

Any pitting in the aluminum layer will be arrested at the thick alloy layer. At the alloy
layer, pits grow in width rather than in depth. The aluminum layer may exhibit abrasion
losses in high velocity rainfall run-of carrying bed load but the alloy layer provides
enhanced resistance to mild-to-moderate abrasion [2]. The alloy layer also resists erosion
corrosion by water and soil, thus providing effective long term protection.

As a consequence of its combined coating properties, aluminized steel achieves a


superior service life over the full range of normal exposure conditions common to
drainage pipe environment. Exceptions include service abrasive conditions and service
corrosive conditions such as exist in seawater, acid mine-water and sanitary sewage [2].
The very slow rate of pit growth observed in the field survey of 30 years old pipes
(Figure 1.6) is due to the galvanic protection mechanism of aluminized steels. The
electro-chemical factors provide effective long term substrate protection throughout the
entire pipe service life [2].
Figure 1.6. Coating Pit on 30-years old pipe demonstrating arrested pipe penetration [2].

The mechanical properties of the coating largely depend upon the thickness of
intermetallic compound layer. Aluminum coatings that contain 5 to 11% Si minimize the
thickness of the iron aluminum alloy bond and improve formability [60].

Above 480ºC further alloying occurs between the aluminum coating and the steel
substrate. Because the rate of alloying is dependant upon time and temperature, all the
coating converts into aluminum-iron-silicon alloy with sufficient time and temperature.
The refractory intermetallic alloy formed is extremely heat resistant and resistant to
spalling upto 680ºC [60]. Spalling at service temperature above 680ºC can be minimized
by the use of heat resistant aluminized steel that contains sufficient Titanium.

Compared with stainless steel, aluminized carbon steels offer greater resistance to attack
by hydrogen sulfide and this resistance is about 100 times more than 18-8 stainless steel
at 595ºC [60].

1.7 Studies on Intermetallic Layer

According to Wang Deqing, Shi Ziyuan and Zou Longjiang [61] the thickness of the pure
aluminum layer on the steel substrate is reduced with the increase in temperature and
time in the initial hot dipping stage, and the thickness of the aluminum layer does not
increase with time at given temperature when identical temperature and complete wetting
occur between liquid aluminum and the substrate surface. The thickness of the Fe–Al
intermetallic layer on the steel base is increased with the increase in hot dipping
temperature and time. The top portion of the coated steel substrate is composed of a thin
layer of α-Al2O3, followed by a thinner layer of FeAl3, and then a much thicker one of
Fe2Al5 on the steel base side.

Shigeaki Kobayashi and Takao Yakou [62] showed that the coating layers of hot dip
aluminized steel consist mainly of aluminum and Fe2Al5. The thickness of the coating
layer increases with increasing immersion temperature. In the diffusion-treatment process
of specimens immersed in molten aluminum, an Fe2Al5 layer preferentially grows and the
coating layer consists of a single phase of Fe2Al5 at temperatures lower than 1273ºK,
while the intermetallic compounds of FeAl and Fe3Al produce at temperatures higher
than 1273ºK. The hardness values of the FeAl and Fe3Al layers are HV600 and HV320,
respectively. These values are lower than the hardness of Fe2Al5 (HV1000) produced at
the normal aluminizing temperature. The fracture resistance of FeAl and Fe3Al are higher
than that of Fe2Al5. The activation energies for the growth of the FeAl and Fe3Al layers
are 180 and 260 kJ/mol, respectively.

According to Sung-Ha Hwang et al [63] the dissolution rate of substrate decreases as the
carbon concentration of the steel specimen increases. An intermetallic layer forms on the
steel substrate as a result of a reaction between the molten aluminum alloy and the steel
substrate. The thickness x of the intermetallic layer increases with dipping time t
following the relationship: x = kt0.5, where k is a growth rate constant. The thickness
decreases as the carbon content of the steel substrate increases. The interface between the
intermetallic layer and the steel substrate also becomes smoother with the increase of the
carbon content. They [63] also suggested a diffusion mechanism controlling the
dissolution of the steel substrate that the carbon atoms in the intermetallic layer reduce
the diffusion rate of iron and aluminum through the layer.
Li Yajiang, Wang Juan, Zhang Yonglan and X Holly [64] observed that the Fe–Al alloy
layer is composed of α-Fe (Al) solid solution, Fe3Al (13.87% Al), FeAl (32.57% Al),
Fe2Al5 (54.71% Al) with different proportions. There is no brittle phase containing higher
aluminum content such as FeAl3 (59.18% Al) and Fe2Al7 (62.93% Al).

Chaur-Jeng Wang and Shih-Ming Chen [65] investigated the high-temperature oxidation
behavior of low carbon steel with hot-dip aluminizing coating in the temperature range
from 750 to 950ºC in air for various durations of time. They found the following results:
(a) The thickness of aluminide layer remained almost the same for all durations of test.
However, the Fe2Al5+FeAl2 dominated aluminide layer transformed to FeAl gradually
with increasing time owing to the aluminum dilution. (b) The accumulation of voids at
the interface between the aluminide layer and the steel substrate might produce cracks
and result in the degradation of the aluminide layer. Therefore, protrusion of iron oxide
nodules on the surface of coated specimen occurred after oxidation at 850ºC for 24 hrs,
(c) Severe cracking and internal oxidation occur in the coated specimen after high-
temperature oxidation. Thus, these limitations degraded the use of aluminized coatings in
high-temperature oxidation environments.

H.R. Shahverdi, M.R. Ghomashchi, S. Shabestari and J. Hejazi [66] examined the
microstructure of solid iron and molten aluminum couple at the interface to identify the
phases as Fe2A15 and FeA13. The former grows faster and is the major phase while the
latter is easily detectable at the later stages of reaction and dissolves partially within
molten aluminum. This is in contradiction with thermodynamic principles, where FeA13
is expected to form before Fe2A15, which may suggest the importance of kinetics in this
system. It was further shown that it is possible to miss FeA13 detection due to the
following reasons:
a) Small quantity of FeA13 due to faster growth rate of Fe2A15, which is directly
related to its defects crystallography.
b) Similarity of XRD spectra of both FeA13 and Fe2A15.
c) Spalling, subsequent re-melting and dissolution of FeA13 into molten aluminum
with time.
d) As reaction proceeds, FeA13 patches join together to form a continuous layer over
the Fe2A15 phase and even after 3000s, its thickness does not exceed more than
10 μm.

1.8 Control of Intermetallic at interface

According to Sung-Ha Hwang et al [63] when steel is in contact with a molten aluminum
alloy, the intermetallic layer grows and dissolves concurrently into the molten aluminum
alloy, which is directly associated with the loss of the steel substrate. The growth and the
dissolution rates of the intermetallic layer determine thickness of the layer. The rates are
closely related with temperature of the molten aluminum alloy and dipping time of the
steel [54,55,56-58]. The thickness of the layer also varies depending on chemical
composition of the molten alloy [59]. A typical alloying element that is known to alter the
thickness significantly is silicon. It is well known that the thickness of the intermetallic
layer decreases with the increase of the silicon content in the melt [54,67,56,59]. During
the HDA process, silicon is normally added into the melt to reduce the thickness of the
brittle intermetallic layer as well as to increase surface hardness of the steel sheet.
However, a limit of silicon concentration in the melt is reported; the layer thickness
hardly decreases with addition of silicon over around 10 wt. % [54,68]. Moreover, the
growth rate of the intermetallic compound layers decreases with increasing carbon
content in the steel substrate [69], and is inhibited by silicon atoms [70,71].

1.8.1 Effect of aluminizing temperature

Shigeaki Kobayashi and Takao Yakou [62] investigated the effect of temperature on the
growth of intemetallic layer between steel substrate and surface aluminum coat. They hot
dip aluminized steel samples at 973ºK, 1023ºK, 1073ºK, 1123ºK and 1173ºK for 300 s.
Figures 1.7 (a-e) show micrographs of cross sections of the coating layers of steel
substrates for these samples. They [62] showed that all the coating layers consist of two
phases. The dark layer at the interface between the layers shown in Figure 1.7(a) was a
shadow from the step in the coating produced by the mechanical polishing process. Their
XRD results showed that the surface and lower layers were aluminum and Fe2Al5,
respectively. These results are consistent with the literature [37,72-75]. Moreover, in the
case of specimens immersed at temperatures higher than 1073ºK, the columnar grains
grew from the Fe2Al5 layer towards the surface aluminum layer. These columnar grains
coarsened and increased with increasing immersion temperature. It has been suggested
that the columnar grains are FeAl3 [37,72-75].

They [62] also showed that the thickness of the coating layer increased with increasing
immersion temperature. With regard to the thickness of the two formed layers, the
thickness of the surface aluminum layer gradually increased from 100 to 200 mm. On the
other hand, the Fe2Al5 layer grows rapidly with increasing immersion temperature at
temperatures ranging from 973ºK to 1073ºK and the growth was saturated at
temperatures higher than 1073ºK.

The interface between the Fe2Al5 layer and steel substrate [62] was irregular due to the
tongue-like morphology of the Fe2Al5 layer. The irregularity of the thickness of the
Fe2Al5 layer increases with increasing immersion temperature. Conversely, the interface
between the surface aluminum and Fe2Al5 layers was flat regardless of immersion
temperature.
Cont…..
Figure 1.7 (a-e) Cross section micrographs of specimens immersed into
molten aluminum for 300 seconds at (a) 973ºK, (b) 1023ºK, (c) 1073ºK,
(d) 1123ºK and (e) 1173ºK [62].

1.8.2 Effect of Diffusion Treatment

Shigeaki Kobayashi and Takao Yakou [62] also discussed the effect of diffusion

treatment on the growth of intemetallic layer between steel substrate and surface

aluminum coat. They diffusion treated five steel samples (HDA at 973ºK, 1023ºK,

1073ºK, 1123ºK and 1173ºK for 300 s) at 873ºK (0.48 TmFe, where TmFe is the

homologous temperature of the melting point of Fe), 1073ºK (0.59 TmFe), 1273ºK (0.70

TmFe) and 1323ºK(0.73 TmFe) for 1.2 ks after immersion and then in another diffusion

treatment for longer time 3.6 ks at the same temperatures. Figure 1.8 shows micrographs

of the cross section of the coating when diffusion-treated for 1.2 ks after immersion.

Figure 1.9 shows the micrographs in the case of a longer treatment time of 3.6 ks. The

intermetallic compound layers in the surface upper layer are denoted by A, B and C. The

black area in the coating layers formed at 1073-1323ºK are likely to be oxides formed

from the specimen surface by high temperature oxidation. Figures 1.10 (a), (b) and (c)

show the results of XRD measurements for layers A, B and C, respectively. Layer A

exhibited an intense peak attributable to (002) Fe2Al5. The individual tongue-like grains
of Fe2Al5 were single crystals and grew in a unidirectional manner perpendicular to the

surface of the specimen. Layers B and C showed peaks of FeAl and Fe3Al, respectively.

These intermetallic layers only grew at relatively higher temperatures than 1273ºK (0.70

TmFe).

It was shown [62] that in the case of the diffusion-treatment at temperatures lower than

the melting point of aluminum, the morphology and thickness of the coating layers hardly

changed from that of the initial coating layers after immersion shown in Figure 1.7 even

though the samples were diffusion-treated for 3.6 ks.

In specimens diffusion-treated at temperatures higher than 1073ºK (0.50 TmFe), the


Fe2Al5 layer (layer A) preferentially grew towards the aluminum layer. In specimens with
a thin aluminum layer formed by immersion at temperatures ranging from 973 to 1023ºK,
the aluminum layer disappeared completely, and the coating layer changed into a layer
consisting of a single phase of Fe2Al5 as shown in Figure 1.8. Aluminum layers
disappeared completely after 3.6 ks of heat-treatment at these temperatures, even in
specimens initially coated with a thick aluminum layer as shown in Figure 1.9. Although
the growth of FeAl3-columnar grains was simultaneous with the growth of the Fe2Al5
layer, the coating layer developed into a single phase of Fe2Al5 due to the preferential
growth of the Fe2Al5 layer.

Moreover, in specimens diffusion-treated at temperatures greater than 1273ºK, FeAl


(layer B) and Fe3Al (layer C) layers were newly formed at the interface between the steel
substrate and the Fe2Al5 layer. FeAl and Fe3Al layers grew with increasing temperature
and time of diffusion treatment. These layers initially exhibited an irregular shape that
reflected the tongue like structure of Fe2Al5 layer in the pre-diffusion treated specimen,
before developing a flat interface with increasing heating time.
As mentioned above, it was found [62] that Fe-Al intermetallic compound layers formed
by hot dip coating and subsequent diffusion treatment transform from layers with a
higher iron composition into layers with a lower iron composition towards the surface of
the specimen. No FeAl2 layers were formed under the conditions examined in this study,
although an FeAl2 phase was formed in chemical compositions ranging from 66 to 67
at. % Al [3].

In the case of diffusion-treatment in air at temperatures higher than the melting point of
aluminum, high temperature oxidation occurred on the surface of the specimen.
Particularly, in pre-diffusion treated specimens with a thin coating layer, the coating layer
peeled away from the FeAl/Fe3Al interface (B/C interface) due to the presence of oxides
at the interface. While oxides preferentially exist in the surface Fe2Al5 layer (layer A) and
the FeAl/Fe3Al interface (B/C interface), the inner FeAl and Fe3Al layers have sound
structures without oxidation.
Figure 1.8. Cross section micrographs of specimens diffusion-treated for 1.2 ks at
873ºK, 1073ºK, 1273ºK and 1323ºK [62].
Figure 1.9. Cross section micrographs of specimens diffusion-treated for 3.6 ks at
873ºK, 1073ºK, 1273ºK and 1323ºK [62].
Figure 1.10. XRD patterns taken from intermetallic compound layers after the
diffusion treatment [62].

1.8.3 Effect of the Carbon Contents of steel substrate

Sung-Ha Hwang, Jin-Hwa Song and Yong-Suk Kim [63] investigated the effect of
carbon contents of the steel substrate on the control of interlayer between steel substrate
and surface aluminum coat. They selected three different steel compositions with
C = 0.202, 0.456 and 1.103 % and then hot dip aluminized in molten Al–9.08 wt.% Si–
0.98 wt.% Fe alloy at 660ºC for 10 min under a rotating condition with two different
speeds of revolution, 16 and 26 rpm. Figure 1.11 shows variation of the dissolved amount
(weight loss) as a function of carbon concentration of these steel specimens. The
dissolved amount decreases with the increase of the carbon concentration at both
revolution speeds. Specimens tested at the higher rotation speed experienced more rapid
dissolution.

Figure 1.11. Weight loss per unit area vs. carbon concentration of specimens hot-
dipped in the molten Al–Si–Fe alloy at 660ºC for 10 min. Specimens were rotated in
the melt with two different speeds of revolution [63].

They [63] also measured the weight loss of the specimens as a function of dipping time,
and the result is displayed in Figure 1.12. The dissolved amount per unit exposed area of
all specimen increases linearly with time. The specimen with C = 0.202% has the lowest
carbon content lost the largest amount, and showed the highest dissolution rate (Figure
1.12). The gradient of plots in Figure 1.12, which represents the dissolution rate,
decreases as the carbon content increases. Calculated gradients of the plots for the
specimens with C = 0.202, 0.456 and 1.103 %, in Figure 1.12 are 1.90 × 10−7, 1.17 ×
10−7, and 1.04 × 10−7 g/mm2, respectively. Figures 1.11 and 1.12 demonstrate that the
carbon content of steel directly influences the dissolution of the steel into the molten
aluminum alloy. As the carbon content of steel increases, less amount of the steel was
dissolved into the molten aluminum alloy with the decrease of the dissolution rate.

0.2 wt% C
0.45 wt% C
1.10 wt% C

Figure 1.12. Weight loss per unit area vs. dipping time for steel specimens with
varying carbon contents, hot-dipped in the molten Al–Si–Fe alloy at 660ºC [63].

Sung-Ha Hwang, Jin-Hwa Song and Yong-Suk Kim [63] also studied the cross sections
of the aluminized steel specimens which revealed intermetallic layers formed between the
top aluminum alloy coat and the steel substrate (Figure 1.13). Intermetallic layers formed
during aluminizing of steel in an aluminum–silicon melt are known to consist of
numerous AlFeSi phases [55,57,76]. The intermetallic layers shown in Figure 1.13 are, in
appearance, composed of two layers, the upper and lower layer. Figure 1.13 shows that
thickness of the intermetallic layer decreases with the increase of the carbon content of
the steel substrate. The thickness of the upper intermetallic layer, which is adjacent to the
aluminum alloy topcoat, does not vary noticeably with the carbon content variation. On
the other hand, the thickness of the lower intermediate layer that is adjacent to the steel
substrate decreases significantly with the carbon content increase. The interface between
the lower intermetallic layer and the steel substrate also becomes smoother as the carbon
concentration increases. These observations are consistent with the previously reported
findings that the composition of steel has more influence on the thickness and interface
morphology of the lower intermetallic layer than on those of the upper layer [55,77].

Figure 1.14 shows variation of intermetallic layer thickness of specimens with different
carbon content as a function of dipping time [63]. The thickness increases with dipping
time for all specimens. The intermetallic layer of the specimen with C = 0.202% grew
much faster than those of the other specimens. The thickness versus time plots in Figure
1.14 look very similar to the weight loss versus time plots in Figure 1.12. Both of the
intermetallic layer thickness and the dissolved amount increased with dipping time, and
both decreased with the increase of carbon content of the steel substrate.
0.2 wt% C

Cont…..

0.45 wt% C
1.10 wt% C

Figure 1.13. Micrographs of intermetallic layers formed on steel specimens hot dipped
in the molten Al–Si–Fe alloy at 660ºC for 40 min, rotated at the speed of 16 rpm:
(a) 0.2%C, (b) 0.45%C, and (c) 1.10%C [63].
0.2 wt% C
0.45 wt% C
1.10 wt% C

Figure 1.14. Intermetallic layer thickness vs. dipping time for steel specimens with
varying carbon content, hot-dipped in the molten Al–Si–Fe alloy at 660ºC. Specimens
were rotated in the melt with the rotation speed of 16 rpm [63].

1.8.4 Role of Silicon

It is well known [59,37,78,79], that after hot dip aluminizing of steel the coating obtained
is of a duplex nature. It contains a brittle layer of intermetallic compounds adjacent to the
steel substrate, which is called the alloy layer. On this a layer of the melt material is
superimposed, which gives good corrosion resistance. As has been discussed in the
literature [59,78], thick alloy layers grow under normal operating conditions in pure and
iron containing aluminum melts. These cannot withstand bending and forming operations
without peeling and flaking. It was pointed out by Gittings et al. [59], that silicon as an
addition to the aluminum melt reduces the thickness of the alloy layer. This can be seen
in Figure 1.15, for the low alloyed steel used in this work after hot dipping in a pure and
silicon containing aluminum melt. Silicon-containing aluminum melts are often used in
commercial hot dip aluminizing.
100 μm

100 μm

Figure 1.15. Alloy layers after hot dipping in pure and silicon-containing aluminum
melts. For 900 sec (a) pure aluminum, 780oC, (b) AI + 2wt% Si, 791oC [59].

1.8.4.1 How silicon could reduce the alloy layer thickness during hot dip aluminizing

There are different explanations for the effect of silicon: Nicholls [80] assumes, that
silicon atoms occupy the structural vacancies of the ε(Al5Fe2)-phase, that has good
diffusion possibilities for aluminum atoms in the pure (binary) state, as was discovered
by Heumann and Dittrich [73]. Lainer and Kurakin [81] report, that in silicon-containing
aluminum melts intermetallic phases (Fex Aly Siz) are obtained, growing more slowly
than the q-phase. In both instances, the presence of silicon results in a slower solid state
growth. Other authors [82,83] take the view, that silicon accelerates the velocity of the
iron enrichment in initially iron-free aluminum melts. If one assumes that the diffusion of
iron atoms through a boundary layer of thickness δ is rate determining, then the velocity
of this iron enrichment is given by

dc/dt = [(k1s)/V](cs - c) ------- (1)

where dc/dt is the velocity of the iron enrichment, k1 is the rate constant of the iron
enrichment, S is the surface of the hot dipped specimen, V is the volume of the melt, cs is
the solubility of the iron in the melt, c is the actual concentration of the iron in the melt.

K1 is given by K1 = D/δ -------- (2)

where k1 is the rate constant of the iron enrichment, D is the diffusion coefficient of the
iron atoms in the boundary layer, δ is the thickness of the boundary layer. It was argued
[83], that silicon could accelerate the iron enrichment, because it increases the solubility
cs (Equation 1) of the melt, as follows from the A1-Fe-Si system reported by Rivlin and
Raynor [84]. Moreover, since the viscosity of silicon-containing aluminum melts is
smaller than the viscosity of pure aluminum melts, smaller δ-values (Equation 2) are
obtained. Komatsu et al. [82] made weight loss studies and report that in the case of
silicon-containing aluminum melts larger weight losses can be observed than in the case
of silicon free aluminum melts. From this they concluded, that silicon acts by
accelerating the iron dissolution.

According to G. Eggeler et al. [56] silicon acts in the solid state when reducing the alloy
layer thickness. When using aluminum melts containing 2 wt % silicon instead of pure
aluminum melts for hot dip aluminizing, the thickness of the alloy layer growing is much
smaller. The silicon is incorporated in the alloy layer, and thus slows its growth. It does
not accelerate the process of iron dissolution in the aluminum melt, opposite to what has
been previously assumed in the literature [82, 83].

M. Vedat Akdeniz, Amdulla O. Mekhrabov and Turgay Yilmaz [71] showed that the
presence of Si in an Fe-Al system decreases the diffusion rate of Al atoms in the Fe
matrix leading to reducing the rate of formation, and hence thickening of the brittle
intermetallic layer formed at the base metal-coat-interface.

According to M. Vedat Akdeniz, Amdulla O. Mekhrabov and Turgay Yilmaz [71] the
formation and the development of interlayer appears to be controlled by the preferred
diffusion of iron atoms into Al-Si coating alloy when Si is added to the coating bath. This
leads to the formation of complex Fex Siy AlZ ternary compounds at the interface. The rate
of growth of such an interlayer is high during a short initial period of immersion but tends
to decrease at a later stage of coating, Figure 1.16. The growth rate of binary Fe-Al
intermetallics is higher than that of interfacial Fex Siy AlZ alloys and the thickness such a
layer is not affected significantly by the temperature, Figure 1.17. There is a rapid
decrease in the thickness of the interfacial layer as the silicon content increases to ~ 2.5
at% Si, Figure 1.18. A smaller decrease occurs as the silicon content further increases.
Figure 1.16. Effect of immersion time on the thickness of the
diffusion layer (T = 800ºC) [71].
Figure1.17. Effect of temperature of the coating bath on the thickness
of the diffusion layer (t = 2 min) [71].

Figure 1.18. Effect of Si content on the thickness of the diffusion


layer (T = 800ºC, t = 2 min) [71].

1.8.5 Role of alloying elements

M. V. Akdeniz and A. O. Mekhrabov [85] calculated the pairwise inter-atomic potential


and ordering energies to predict the effect of various alloying additions on the activity co-
efficient of Al-atoms in α-Fe0.95(Al1-nXn)0.05 alloys where X =Si, Ti, Ge, Sb, Mg, Cu, Ca,
Ag, Cd, Cr, Co, Zn, Mn, Ni, Pb or Bi considered upto 1 at%. The results of calculations
showed that the impurity elements with regard to their effect on the activity co-efficient
of Al-atoms may be classified into two groups; I-group impurity elements, X1 = Si, Ti,
Ge, Sb, Mg, Cu, Ca, Ag, Cd and Cr which decrease the activity co-efficient of Al-atoms,
reduce the thickness of Fe-Al intermetallic diffusion layer and II-group impurity elements
XII = Co, Zn, Mn, Ni, Pb and Bi, which increase the activity co-efficient of Al-atoms tend
to increase the thickness of diffusion layer at the Fe-Al interface
H. Glasbrenner and E. Nold, Z. Voss [86] doped the Al-melt with one of the elements
Mo, W or Nb with a nominal composition of about 1 wt% during HDA. They studied the
influence of these elements on the coating formed and on the following oxidation process
Hot dip aluminizing was carried out at 800°C for 5 min under dry Ar atmosphere. The
oxidation experiments were performed at 950°C for 24 h in air. Compared to the HDA
processes with pure A1, the addition of the alloying elements leads to a thinner
intermetallic layer.

2 EXPERIMENTAL

2.1 Size, Shape and Chemical Composition of Steel Specimens

Four different grades of commercial plane carbon steels were used in the present
experiments. The actual compositions of the steels used are given in Table 2.1.

Table 2.1. Chemical composition of the steels used in the present study.

Carbon Manganese Silicon Sulphur Phosphorous


Steel # 1 0.13 % 0.68 0.16 0.027 0.008

Steel # 2 0.25 % 0.72 0.19 0.022 0.008

Steel # 3 0.45 % 0.72 0.18 0.030 0.011


Steel # 4 0.60% 0.70 0.19 0.025 0.012

Three different types of specimens were used:

 Cylindrical specimens of 10 mm diameter and 30 mm length,

 Flat specimens measuring 40x20x2 mm

 Flat specimens measuring 100x20x2 mm

Most of the experiments were conducted on flat specimens, so that some of these
specimens could be subjected to X-ray diffraction, and also to a ‘bend’ type test which
was designed to compare the relative brittle behavior of the various coatings. A hole of
~2 mm diameter was drilled at one end of the specimen to facilitate its hanging into the
molten aluminum. The shapes of specimens are illustrated in Figures 2.1 a & b.

(b)

(a)
Dia = 10 mm Size = 100x20x2 mm
Length = 30 mm

Figure 2.1. The shapes of the specimens used in aluminizing experiments.

2.2 Preparation of Aluminum Melts

Some of the initial experiments were conducted in an electrically heated pit-type furnace.
However, it was observed that during the aluminizing of 100 mm long specimens, it was
difficult to maintain the temperature over the entire length of the specimens. Hence, a
gas-fired furnace with a double burner was used. The design of the furnace is shown in
Figure 2.2.

The aluminum alloy was melted in a graphite crucible in a gas fired-furnace in which the
(a)
(b)
temperature could be controlled to within ±5°C with the help of a K-type thermocouple /
controller. The molten aluminum alloy was covered with a layer of KCl-NaCl-Na2F flux.
The ratio of the constituents was adjusted to achieve a suitable melting point of the flux.
Before every coating experiment the temperature was carefully measured and controlled
at the required level. After hot-dipping of the samples for the required period, the samples
were taken out and quenched into boiling water. In some experiments, the samples were
lightly centrifuged before quenching.
Figure 2.2. Schematic diagram of the gas fired furnace. The burner A was a high-heat
gas blower type burner, controlled through an on/off K-type temperature controller.
The burner B was a small burner which was kept burning at a low-heat all the time
during experiments. This arrangement was helpful in maintaining the temperature
within a narrow range.

Aluminizing in pure aluminum was carried out in 100 % aluminum melt prepared by
melting electrical grade conductor of aluminum that was available in small cut-pieces.

The additions of silicon were made by adding calculated quantities of Al-50% Si master
alloy. The actual composition of the melt was also obtained by taking small samples from
the molten alloy which was later analyzed on a ‘METOREX’ ArcMet-930 Spark
Emission Spectrometer.

2.3 Cleaning / Fluxing prior to Aluminizing

Three different cleaning methods were employed during the present work:
1. In the first method, the samples were pickled with HCl, rinsed in running water,
treated with an alkaline solution, and then dipped in molten flux for 2-3 minutes
before transferring into the aluminizing melt. The flux consisted of a mixture of
NaCl, KCl, and Na2F, in which the ratio of the three constituents was adjusted to
achieve a suitable melting point. However, most of the experiments were carried out
in a flux of the following composition:

NaCl = 4 parts
KCl = 5 parts
Na2F = 1 part

2. In the second method the pickled samples were:

 dipped for 1-2 minutes, in a near-saturated aqueous solution of the flux kept at 80-
90 degree C,

 dried in still air,

 and then immediately transferred into the aluminizing melt.

3. Most of the experiments were however carried out by this third method. In this case
the mechanically cleaned samples were carefully treated / pickled in a 50 % diluted
solution of HCl, rinsed in running water and then in methanol, and then dried with a
hot-air blower. The samples were treated in this manner and then immediately dipped
into the aluminizing bath.

The method outline at # 3 provided an easy and yet very effective method of cleaning for
the present purpose. However, it must be realized that whereas this method can be safely
used for carefully handled individual research specimens, the method was clearly not
easy to follow in a commercial aluminizing process. When accomplished successfully,
the results of this method were ‘good’ as illustrated by Figure 2.3.

(a)

(b)
Figure 2.3. Examples of (a) an improperly cleaned sample, and (b) a properly
cleaned sample.

2.4 Bending of Aluminized Specimens

Flat aluminized samples were subjected to a bend type test, in order to compare the
relative ‘ductility’ of the samples aluminized in various aluminum melts. A bending
machine which was especially designed for this purpose is shown in Figure 2.4. The
methodology of the bend test and the details of the machine are outlined in Chapter 3.
Figure 2.4. Bending machine used for conducting bend type
test of hot-dip aluminized samples.

2.5 Metallographic Examination of Aluminized Specimens

The aluminized samples were suitably sectioned, mounted in bakelite, and then
metallographically prepared before etching in 2% Nital solution. Microscopic
examination of the polished and etched samples was carried out on an Olympus GX-1
camera microscope. This microscope was fitted with a ‘Moticam’ digital camera coupled
with a computer. The various magnifications were calibrated with the help of a stage
micrometer.
Specimens subjected to bend test were also studied metallographically to examine the
behavior of interlayer during bending. For this purpose, the samples were sectioned
across the line of the cracking observed on the specimen surface.

2.6 X-ray Diffraction of Aluminized Samples

X-ray diffraction of the selected samples was carried out, on a Panalytical Xpert-Pro
diffractometer, to identify the phases formed in the interlayer. For this purpose, flat
samples were studied in such a manner that after obtaining an XRD pattern, an
approximately 5-10 micron surface layer was carefully removed by grinding on a 600 grit
emery paper, and an XRD pattern taken again. This process was repeated till the entire
aluminum layer and then the interlayer were completely removed, and the final XRD
pattern indicated mainly the substrate steel being present.

2.7 Scanning Electron Microscopic Examination

Selected aluminized samples were studied on a Hitachi S-3000H scanning electron


microscope, which is fitted with an EMAX X-ray detector. Individual phases were also
analyzed for their chemical composition.

2.8 Tensile Testing of Al-Si alloy samples

Selected samples taken from aluminum-silicon melts used during the present
experiments, were also subjected to tensile testing. Fractured tensile specimens were
studied with optical as well as scanning electron microscope, to examination the fracture
mechanism.
3 RESULTS AND DISCUSSION

3.1 Aluminizing in pure aluminum

Most of the samples which were aluminized for microstructural and XRD examination
were the 100x20x2mm size. As explained in the later sections, these specimens were also
subjected to a bend type test. Photographs of some typical aluminized samples are given
in Figure 3.1, to illustrate the kind of surface finish that was obtainable during the present
work. A complete dipping of the specimens was not preferred because in such a case the
molten aluminum, which flowed out of the ‘hole’ when the specimen was drawn out of
the melt, created an un-even surface layer on the aluminized samples. A typical sample
on which the flown aluminum had affected the surface is shown in Figure 3.2.

Microstructure of hot-dip aluminized samples of low carbon steels, for two different
dipping times, are shown in Figure 3.3. Three distinct regions which could be easily
identified in these microstructures include: the outer aluminum layer, the inter-metallic
compound layer, and the substrate steel. A typical feature of aluminizing in pure
aluminum (in the temperature range of 700-750ºC) was the un-even interface between the
intermetallic compound layer and the substrate steel. It can be seen that the growth of the
compound layer into the steel occurs in a ‘finger-like’ manner similar to the ‘dendritic’
growth during solidification, i.e., individual ‘fingers’ of the compound grow into the steel
like the primary dendrites.

It was noted that the details inside the interlayer were not revealed by the conventional
Nital etching. On the other hand, etching in an aqueous solution of hydrofluoric acid was
able to reveal the microstructure inside the interlayer, but was not able to etch the steel
substrate. Thus, during the present work a ‘double etching’, i.e., first with Nital and then
with HF was used whenever it was needed to examine both the interlayer as well as the
substrate in the same sample. Microstructures shown in Figure 3.4 illustrate the effect of
single and double etching.
samples taken for
XRD and bend test

(a)

(b)

Figure 3.1. Photographs of aluminized samples illustrating the typical degree of


the surface finish obtained during the present experiments.
flown
flown
aluminum
aluminum

Figure 3.2. Photographs of typical samples on which the flown aluminum had
affected the quality of the aluminized surface.

A closer examination of the microstructures of Figures 3.3 and 3.4 reveal that the growth
of ‘fingers’ into the steel appears to be hampered by the pearlitic areas in the
microstructure. Two important observations that may be made from the microstructures
of Figures 3.3 & 3.4 are:
 The compound layer, during its advance into the steel tends to grow through the
ferrite phase, by-passing the pearlitic areas, with the result that the pearlitic
regions get enveloped / entrapped into the compound layer. The entrapment of
pearlite (or carbides) can be more clearly observed in the microstructure of Figure
3.3 (b).
 As the compound phase advances into the steel, the carbon in the steel diffuses
ahead of the advancing interface causing a ‘carbon build-up’, which is evidenced
by an increased volume fraction of pearlite phase ahead of the transformation
front.
Figure 3.3. Microstructure of a hot-dip-aluminized steel showing the outer
aluminum coat, the intermediate compound layer, and the substrate steel, for (a) 1
min at 700°C, and (b) 3 min at 750°C. The difference in the thickness of the
compound layers may also be noted. The accumulation of pearlite ahead of the
interlayer’s advancing front may also be noted in Figure 3.3 (b).

(a)

carbon
build-up

25 μm etched in Nital

(b)

Entrapped
carbide particles

etched in Nital
25 μm and then in HF
Figure 3.4. (a) Microstructures of hot dip aluminized sample etched in Nital.
(b) Microstructure of the sample etched in Nital and then in HF. The effect of double
etching in (b) may be noted.

From the carbon build-up as observed in Figure 3.3 (b), it may be concluded that the
carbon should have an extremely limited solubility in the Fe-Al intermediate compound.
It should also follow from this observation that with an increase in the carbon content of
the steel (i.e., a greater volume fraction of pearlite in the microstructure), the rate of the
growth of interlayer into the steel may be expected to decrease.

(a)
(b)

Figure 3.5. The thickness loss Vs dipping times of 10, 20 and 30 minutes at 700ºC for
low and medium carbon steel samples in (a) pure aluminum, and (b) in Al-6%Si alloy.

This inference was confirmed from the results of an experiment carried out during the
present work in which steel samples of different carbon contents were treated in molten
aluminum, and their dissolution rates compared. Figure 3.5 summarizes the results of this
experiment in which steel samples of equal thickness and different carbon contents were
immersed into molten aluminum at 700ºC for upto 30 minutes after which their thickness
as well as weight loss were measured. The results clearly suggested that the dissolution
rate decreases with an increase in the carbon contents. This result is consistent with the
work of Sung-Ha Hwang et al [63] in which they have reported a similar observation.

Another interesting consequence of the ‘carbon build-up’ ahead of the transformation


front was noticed in low-carbon steel samples which were aluminized at a temperature of
750°C and were quenched into water. In these samples, the martensite phase was
observed to have formed in the regions of the ‘carbon build-up’. Microstructures
illustrating this effect are shown in Figure 3.6. Clearly, these regions were composed of
austenite at the aluminizing temperature of 750°C, and had transformed into martensite
upon cooling following the aluminizing. It may be noted that the microstructure at the
inner portion of the steel did not transform into martensite, owing to the ‘low’ carbon
concentration at the interior. In fact, the martensitic regions, like the one shown in Figure
3.6, were only observed in those samples which were promptly quenched into water after
the aluminizing.

Various workers have used temperatures exceeding 750°C for hot dip aluminizing
[71,75] but none of them seem to have considered that the aluminizing reaction that they
were studying was actually taking place between aluminum and austenite and not
between aluminum and ferrite plus pearlite.

An important inference that can be drawn from the microstructure of Figure 3.6 is that
even when the microstructure ahead of the interface consisted of only the austenite phase
(i.e., when there was no pearlite phase present) the growth pattern of the interlayer was
still finger-like. This in turn means that the finger-like growth is not caused by the
presence of the pearlitic regions in the microstructure, but instead it is an inherent
characteristic of aluminizing with pure aluminum. This inference is also supported by
Figure 3.3 (a) in which a finger-like growth can be seen to develop even on very short
dipping times when no interference by the pearlitic regions can yet be evidenced.

The above observations do not support the work of Sung-Ha Hwang et al [63] in which
they concluded that the finger-like (which they called as tongue-like) growth mode was
actually resulted from the interference of the pearlite phase with the growing interface.
The present work clearly shows that although the growth of the interlayer may indeed be
retarded by the pearlite phase in the microstructure, the finger-like morphology is not
related to the interference with pearlite. The ‘fingers’ can be seen to form even in low
carbon steels on very short dipping times (i.e., when no interference with pearlite could
begin), and also, when a single-phased austenitic region may be present ahead of the
advancing interface. Nevertheless, it is clear that the interference with the pearlitic
regions do have an impeding affect on the advance of the ‘fingers’.

In this connection an interesting experiments was conducted in which a 0.6 % carbon


steel that had been quenched and tempered to obtain a uniform distribution of
spheriodized carbides was aluminized. The idea behind this experiment was to examine
the interference of a uniform distribution of carbides with the advance of the ‘fingers’.
Microstructure given in Figure 3.7 shows that although the finger-like character can still
be evidenced, the interface between the intermetallic and the substrate is considerably
smoother than that seen in Figure 3.6. An obvious conclusion that can be drawn from this
observation is that the interference with the pearlite or carbides in the microstructure of
the substrate can affect the profile of the interface and thereby, possibly, the rate of its
advance as well. The effect of the volume fraction of carbides in the microstructure of the
steel on the growth rate of the interlayer has also been reported in the work of Sung-Ha
Hwang et al [63].

(a)
(b)

martensite

25 μm

Figure 3.6 (a & b). Microstructures of the samples hot-dip-aluminized at 750°C, and
water-quenched, showing the formation of martensite in the region of carbon build-up
ahead of the advancing interface. It may be noted that the growth front still exhibits a
finger-like morphology.
carbon
build-up

25 μm

Figure 3.7. Microstructure of a HDA spheriodized steel sample. A considerably


smooth interface between intermetallic and steel substrate can be seen in this
microstructure, despite the presence of finger like character.

Aluminized samples of medium carbon steel exhibited another feature that has not been
reported anywhere in the published literature. As may be seen in Figure 3.8, a continuous
row of light-etching particles is present ahead of the growing fingers of the interlayer.
Although, it was difficult to establish whether these light-etching particles in the steel
were ferrite or cementite, it appears logical to believe that these particles may be of
ferrite, which is continuously formed ahead of the interface so as to make way for the
intermetallic to grow into the substrate. It must be remembered that the carbides have
little or no solubility in the intermetallic phase. However, during the present work it was
not established what these light-etching particles actually were, or under what conditions
these particles were formed. Further work is required to characterize these particles.
(a)

white etching row


of precipitates

10 μm
(b)

white etching row


of precipitates
8 μm

Figure 3.8 (a & b) Microstructures showing a continuous row of light-etching particles


present ahead of the growing fingers of the interlayer.

To summarize, it may be concluded that when aluminizing is carried out in pure


aluminum, the interlayer grows into the substrate steel with a finger-like morphology.
However, the profile of the interface and even the growth-rate of the interlayer can be
significantly affected by the amount and the distribution of the pearlite in the
microstructure.
3.2 X-ray Diffraction of samples aluminized in pure aluminum

Samples aluminized in pure aluminum were studied with X-ray diffraction. For this
purpose, flat samples measuring 20x20 mm were studied in such a manner that after
obtaining an XRD pattern, an approximately 5-10 micron surface layer was carefully
removed by emerying on a 600 grit emery paper, and an XRD pattern taken again. The
process was repeated till the entire aluminum layer and then the interlayer were
completely removed, and the final XRD pattern indicated mainly the substrate steel being
present.

Figure 3.9 shows the series of XRD pattern taken from a sample aluminized with pure
aluminum. This sample, after the hot-dipping step, was lightly centrifuged so as to
remove as much aluminum from its surface as possible. The purpose was to be able to
reach the intermetallic layer with minimal grinding. In these XRD patterns, the
diffraction peaks from aluminum (fcc), iron (bcc), and intermetallic compounds have
been marked with suitable symbols.

The first pattern (Figure 3.9 a) which was taken from the as-aluminized surface, i.e.,
without any grinding (Figure 3.9 a) shows all the low-order aluminum peaks alongwith
two additional peaks; one at 43-degree, and the other at 93-degrees. As shown later, the
43 and 93 degrees peaks belonged to the Fe2Al5 phase. In the next few patterns, the
intensity of aluminum peaks gradually decreased while those of the Fe2Al5 phase
gradually increased. It may be noted that in the diffraction pattern of Figure 3.9 (f and g),
the aluminum peaks have almost disappeared and the intensity of the 43-degree
diffraction peak of Fe2Al5 has achieved an exceptionally high value of ~90000 cps.

It is relevant at this point to discuss the two diffraction peaks; at 43-degrees, and 93-
degrees. The diffraction peak at 43-degrees, which corresponds to a d-spacing of ~2.1 Å,
could be indexed as the (002) reflection from Fe2Al5, while the diffraction peak at 93-
degrees which corresponds to (004) reflection, or a second order reflection from the (002)
planes. The corresponding X-ray file and the ‘line-pattern’ from the ICDD are
reproduced in appendix-I.

A comparison of this line-pattern with the diffraction patterns of Figure 3.9 (f and g) shall
suggest that if the diffracting phase is indeed the Fe2Al5 phase, then only the (002) planes
are in the diffracting orientation, as no other peak was present in the diffraction pattern.
And also, that these (002) planes were oriented parallel to the aluminized surface of the
flat specimens.

Further, it was also observed that with successive grinding of the surface the intensity of
this (002) reflection increased to exceptionally high values before starting to decrease
again, suggesting that a very large number of grains/crystal-planes contributed into the
peak intensity at one stage. An obvious explanation of this effect is that the finger-like
crystals exhibited a preferred orientation, i.e., the (002) planes of all the finger-like grains
were oriented parallel to the surface of the specimen, and that all the fingers had the same
growth direction, which was [002].

It must be appreciated that although the 43-degree diffraction peak corresponding to a


d-spacing of ~2.1 Å, could be indexed as the (002) reflection from Fe2Al5, the structure
could not be unambiguously confirmed on the basis of just one diffraction peak. In order
to obtain a confirmation that the finger-like intermediate compound phase was indeed the
Fe2Al5 phase, powder specimens of the interlayer phase were produced by following the
techniques described in the following section. Accordingly, the powder diffraction of the
interlayer gave a diffraction pattern, Figure 3.10, which was consistent with the Fe2Al5
structure. This experiment clearly showed that the finger-like crystals were indeed the
orthorhombic Fe2Al5 phase, and that, these finger-like columnar crystals had grown into
the steel perpendicular to the aluminized surface and along the [002] axis of Fe2Al5, and
thus the (002) planes were oriented normal to the aluminized surface. This observation is
consistent with the work of Heumann and Dittrich [73], although they were not able to obtain
a complete diffraction pattern of the Fe2Al5 interlayer. Later workers [87,74] preferred to
refer to the work of Heumann and Dttrich [73] and believed the finger-like crystals to be
composed of Fe2Al5, without carrying out any detailed diffraction work on their own.

= Al
= Fe2Al5

(a)
= Al
= Fe2Al5

(b)

Cont…..

= Al
= Fe2Al5

(c)
= Al
= Fe2Al5

(d)

Cont…..
= Al
= Fe2Al5

(e)

= Al
= Fe2Al5

(f)

Cont…..
= Al
= Fe2Al5

(g)

= Fe
= Fe2Al5

(h)
Cont…..

= Fe
= Fe2Al5

(i)
= Fe
= Fe2Al5

(j)

Cont…..

= Fe
= Fe2Al5

(k)
= Fe
= Fe2Al5

(l)

Figure 3.9 (a-l) X-ray diffraction patterns obtained from samples aluminized in pure
aluminum. The XRD patterns were taken by successively grinding about 5-10 microns
of the surface layer each time before the next XRD pattern was taken.

(a)
(b)

Figure 3.10. XRD pattern, shown in Figure 3.10 (b), was taken from the powdered
sample of the interlayer obtained from the aluminized sample of Figure 3.11; all the
peaks of the Fe2Al5 phase (which are consistent with the ‘X-ray file’, Appendix-I) can
be seen to be present. For comparison sake, the diffraction pattern obtained from
aluminized flat samples is given in Figure 3.10 (a).

3.3 Obtaining Powdered samples of the interlayer

In order to obtain powdered samples of the interlayer for XRD, thin sheets of low-carbon
steel were aluminized in pure aluminum for prolonged times such that the sheets
appeared to have dissolved almost completely into the molten aluminum. The aim was to
produce samples in which the compound layers had grown from both the opposite
surfaces of the sheet so as to meet each other with very little steel in between. These
samples, when produced successfully, were so brittle that they could easily be crushed
into fine powder. A micrograph taken from such a sample is given in Figure 3.11.
These samples were then crushed in agate mortar to obtain fine powder. Any aluminum
present on the aluminized sheets did not convert into fine powder, but instead remained
as coarse flakes which were then separated so as to leave the fine powder of the inter-
metallic layer only. Any iron particles from the substrate steel were also removed with
the help of a magnet.

The powder diffraction pattern obtained from the fine powder of the intermetallic, shown
in Figure 3.10 could be indexed as the Fe2Al5 phase.

(a)

Cont…..
(b)

100 μm

Figure 3.11 Microstructure of an initially 0.75 mm thick sheet of low carbon steel
aluminized for about 30 minutes. It may be noted that the light-etching regions of the
Fe2Al5 inter-metallic constitute the largest volume fraction in the microstructure. At
this stage the sheet could easily be crushed into fine powder of the inter-metallic.

3.4 Aluminizing in Al-Si alloys

The next series of aluminizing experiments which were conducted on the steel samples
made use of aluminum-silicon alloys of various compositions. Silicon content of upto
13 % were used in these experiments. The microstructures given in Figure 3.12 show the
typical morphology and the thickness of the interlayer formed at two different silicon
contents. A comparison of these microstructures with those of Figure 3.3 & 3.4 reveals
that whereas with hot-dip-aluminizing (HDA) in pure aluminum the interface exhibits a
finger-like or dendritic profile, the HDA with Al-Si alloys yields a more ‘planar’
interface. Further, the thickness of the compound layer in HDA with Al-Si alloys is, in
general, much smaller than that obtained in HDA with pure aluminum.
Microstructures given in Figure 3.12 also indicate that the thickness of the interlayer as
well as the profile of the interlayer/substrate interface are both affected by the actual
silicon content of the melt. It was observed that the interface became almost completely
‘planar’ at a silicon level of about 6 %. Any further increase in the silicon level did not
cause any appreciable change in the thickness or the profile.
Figure 3.12. Microstructures of samples aluminized in (a) Al-1.8%Si, and (b) Al-6%Si
alloys. The difference in the thickness of the interlayer in the two samples may be noted.
It may also be noted in Figure 6a that although the interlayer exhibits a fairly uniform
thickness, the finger-like character can still be seen. Also, in Figure 6b, with 6%Si
addition, the interface between the interlayer and the substrate is very smooth and planar.

Microstructures given in Figure 3.13 show that similar to the aluminizing with pure
aluminum, the carbon build-up ahead of the advancing interface was also observed in the
samples aluminized in Al-Si alloys. However, while the accumulation of the pearlite
ahead of the advancing interlayer was clearly evidenced, no entrapment of pearlite in the
interlayer was noticed. It may further be noted that the accumulated pearlite formed a
continuous band of more or less uniform thickness ahead of the advancing interface.

Another important observation made during the present work was that the finger-like and
the planar morphologies of the interface in the two types of samples were not affected by
the dipping time. This observation strongly suggests that the profile of the interface was
not related to the kinetics of the process, but instead it was most likely dictated either by
the energy of the interface between the substrate steel and the interlayer, or possibly by
the effect of the presence of silicon on the diffusion through the interlayer.
Cont…..

Figure 3.13. Microstructures of samples aluminized in Al-6%Si alloys, for (a) 2


minutes, and (b) 20 minutes. The accumulation of pearlite (carbon build-up)
ahead of the interlayer’s advancing front may be noted.

The above explanation tends to arouse interest in exploring the crystallography and
chemistry of the compound layers formed in the samples aluminized in Al-Si alloys.
Accordingly, selected samples were studied by XRD and the scanning electron
microscopy for the characterization of the interlayers formed in the two types of samples.
The results of the XRD and SEM studies are given in the following section.

3.5 X-ray Diffraction and Scanning Electron Microscopy of samples aluminized in


Al-Si melts

The XRD patterns obtained from the samples aluminized with Al-Si alloys (Figure 3.14)
showed the presence of a phase that could be indexed as Fe3Al0.5Si0.5 [88]. Hereafter, this
phase shall be regarded as the Fe3(Al,Si)1 or even the Fe3Al phase. The Fe3(Al,Si)1 phase
has the cubic-DO3 structure [89] with a lattice parameter about twice that of iron. It must
be appreciated that this structure is actually based on Fe3Al compound, and that any
silicon atoms only substitute for the aluminum atoms in the Fe3Al lattice. The unit cell of
the Fe3Al lattice is shown in Figure 3.15, while the respective card from the X-ray file
(ICDD) is reproduced in appendix-I.

= Al
= Fe2Al5
= Fe3Al
= Si

(a)
= Al
= Fe2Al5
= Fe3Al
= Si

(b)

Cont…..

= Al
= Fe2Al5
= Fe3Al
= Si

(c)
= Al
= Fe2Al5
= Fe3Al

(d)

Cont…..
= Al
= Fe2Al5
= Fe3Al
= Fe

(e)

= Fe2Al5
= Fe3Al
= Fe

(f)

Cont…..
= Fe2Al5
= Fe3Al
= Fe

(g)

= Fe

(h)
Figure 3.14 (a-h). X-ray diffraction patterns obtained from the samples aluminized in
Al-6%Si alloy melt. These XRD patterns were taken by successively grinding off
about 5-10 microns of the surface layer each time before the next pattern was taken.

Figure 3.15. Unit cell of the DO3 structure of Fe3(Al,Si) phase [89].

The diffraction patterns shown in Figure 3.14 were obtained from a sample that had been
aluminized in Al-6%Si alloy. These patterns were taken by successively grinding off
about 5-10 microns of the surface layer each time before the next pattern was taken.
The prominent phases that can be identified in these patterns include:
 the FCC-aluminum (in the outer coat),
 the diamond-cubic precipitates of Si (inside the coating),
 the Fe3Al and the Fe2Al5 intermetallics (in the interlayer), and
 the BCC-iron (substrate)

It may be noted in Figure 3.14 that the (002) peak of Fe2Al5 (which was observed in the
samples aluminized in pure aluminum, Figure 3.9) is also present in the XRD patterns
obtained from the samples aluminized in Al-6%Si alloy, which indicates that the
interlayer consisted of two phases; the Fe3Al and the Fe2Al5.

It was of interest to explore the effect of silicon level on the relative volume fractions of
the Fe3Al and the Fe2Al5 phases in the interlayer. Accordingly, flat samples aluminized in
gradually increasing silicon-containing aluminum were studied with XRD. It was
observed that at a silicon level of around 1.5%, the DO3 phase (based on Fe3Al) could be
clearly evidenced in the XRD patterns. With further increase in the silicon-level of the
melt, the relative volume fraction of the DO3 phase gradually increased, as evidenced
from the relative peak-intensities of the Fe3Al and the Fe2Al5 phases in the XRD patterns.
The finger-like interface between the interlayer and the substrate also became
successively smoother, becoming almost ‘planar’ at about 6% silicon, see Figure 3.12.

Although the XRD patterns showed the presence of the Fe3Al and the Fe2Al5 phases, the
distribution of the two phases inside the interlayer could not be seen in the samples
etched in Nital. In an effort to be able to see the relative distributions of the Fe3Al and the
Fe2Al5 phases in the interlayer, some samples were etched in 1% aqueous solution of
hydrofluoric acid, which revealed the microstructural details inside the interlayer. The
microstructures of samples aluminized in Al-1.8%Si and Al-6%Si alloys, Figure 3.16,
show that the interlayer in these samples actually consisted of two layers.

The scanning electron micrographs of Figure 3.17, show the details inside the interlayer
at 0%, 1.8%, 3%, and 6% silicon-levels. The layer next to the substrate (with a finger-like
morphology at low-silicon-levels; believed to be Fe2Al5) gradually reduced in thickness
as the silicon-level was increased, and that, the morphology of this layer successively
changed from finger-like to ‘planar’. The outer layer (adjacent to the outer aluminum-
coating, believed to be Fe3Al) gradually built up in thickness with increasing silicon level
of the melt.

The changes in the thickness of Fe3Al and the Fe2Al5 layers with increasing silicon
additions are summarized in Figure 3.18.
(a)

outer
Al-coat steel
substrate

Fe2Al5
Fe3Al

cracks

20 μm

Cont…..

(b)

Figure 3.16. Optical micrographs of the samples (a) aluminized in Al-1.8%Si alloy,
(b) Al-6% Si alloy showing the two layers of Fe-Al intermetallics in the interlayer,
as revealed by the hydrofluoric acid etching. The cracking in the Fe3Al layer may
also be noted.
(a)

steel
substrate
Fe2Al5

Al-coat

Cont…..

Fe3Al (?)

Fe2Al5

Al-coat

steel
(b) substrate
(c)

Cont…..
(d)

(e)

Fe3Al Fe2Al5

Al-Si alloy coat

Cont…..
(f)

(g) steel
substrate

crack

Fe2Al5
Fe3Al

Al-Si alloy coat


Figure 3.17. Microstructures of the samples aluminized in:
(a & b) pureAl; (c) Al-1.8% Si; (d & e) Al-3% Si; (f & g) Al-6% Si

Microstructures of the samples etched in 1% aqueous solution of hydrofluoric acid. It


may be noted that the Fe3Al phase gets severely attacked by the hydrofluoric acid
etchant; the etching attack may also have been assisted by the cracks/micro-cracks in
the brittle Fe3Al phase. The samples of Figure 3.17 (a, d, e and g) were also etched in
Nital after the etching with hydrofluoric acid, to reveal the microstructure of the
substrate steel.

Figure 3.18 Changes in the thickness of Fe2Al5 and Fe3Al layers with increasing
silicon addition.

Microprobe analyses of the two layers carried out during the present work also showed
the difference in the iron and aluminum concentrations of the two; the layer next to the
substrate (which exhibited finger-like morphology at 0% and 1.8% silicon levels, see
Figure 3.17 a, b and c) was richer in aluminum, while the outer layer (which showed deep
etching with aqueous hydrofluoric acid (see Figure 3.17 c, d, e, f and g), was richer in
iron.
Although the XRD patterns of samples aluminized in pure aluminum (Figure 3.9) did not
indicate the presence of any Fe3Al, the microstructure of Figure 3.17 (a & b) suggests
that small quantities of Fe3Al may be present along the outer boundary of the interlayer,
however, the volume fraction of this Fe3Al may be too small to show up in the diffraction
pattern.

The XRD patterns obtained from the samples aluminized in silicon-containing aluminum
melts provided support to the view that the interlayer was composed of two phases, the
Fe3Al and the Fe2Al5. It was further indicated from the XRD patterns that the Fe2Al5
phase was located adjacent to the substrate; in the XRD patterns obtained from a sample
aluminized in Al-1.8%Si alloy (shown in Figure 3.19), it may be observed that the peak
intensities of the Fe2Al5 phase gradually increased while those of the Fe3Al phase
decreased as the surface of the sample was successively removed by further grinding. It
was also observed that when the finger-like character of the interface disappeared
completely, the XRD patterns showed very small volume fraction of Fe2Al5 to be present.
The microstructures of Figure 3.17 also showed that with an increasing silicon-content,
the thickness of the Fe2Al5 layer gradually decreased while that of the Fe3Al layer
gradually increased.

The presence of Fe3Al phase in the interlayer as observed during the present work is
consistent with the work carried out by L. Yajiang, W. Juan, Z. Yonglan, X. Holly [64]
using transmission microscopy. However, the FeAl3 phase reported by H.R. Shahverdi,
M.R. Ghomashchi, S. Shabestari, J. Hejazi [66] was not observed in any of the samples
studied during the present work.
(a)

Cont…..

(b)
Figure 3.19. X-ray diffraction patterns obtained from samples aluminized in Al-
1.8%Si alloy melt. The XRD pattern of Figure 11a was taken after grinding off the
outer aluminum layer, while the XRD pattern of Figure 11b was taken after a further
grinding of 20-25 microns. The change in the ratio between the peak intensities of
Fe2Al5 and Fe3Al may be noted.

It is important to note that at high silicon levels (6% or above), the Fe2Al5 phase does not
exhibit a finger-like morphology, and also that even when the thickness of the Fe2Al5 has
reduced to only around 1-2 microns, a continuous thin layer is still present along the
interface with the substrate. This observation clearly suggests that:

a) Whether the aluminum-melt contains any silicon or not, the advance of the
interlayer into the steel is governed by the growth of the Fe2Al5 phase; the
microstructures shown in Figures 3.16 and 3.17 clearly show that irrespective
of the silicon level, the phase at the transformation-front is always Fe2Al5.
b) The additions of silicon into the aluminizing melt cause the finger-like growth
mode to gradually disappear. This is believed to be either due to an increase in
the interface energy, or due to a reduction in the diffusion through the Fe2Al5
layer.

3.6 Lattice Parameter Measurements of Fe3Al and Fe2Al5

The measurements of the lattice parameters of the Fe3Al and the Fe2Al5 phases at various
silicon-levels provided interesting results. It was observed that as the silicon level of the
melt was gradually increased, the lattice parameter of the Fe3Al-based cubic phase
showed a continuous decrease, indicating a gradually increasing substitution of aluminum
atoms by silicon on the Fe3Al cubic lattice. This observation was consistent with the
measurement of the silicon-concentration in the interlayer phases carried out during the
present work through the microprobe analysis.
Further, it was also observed that as the silicon level of the melt was gradually increased,
the lattice parameter of the Fe2Al5 phase showed a corresponding increase. This increase
is summarized in Table 3.1. A possible explanation of this effect may be as follows: The
growth of the Fe2Al5-fingers into the steel is governed by the diffusion of aluminum and
iron atoms through the interlayer. Keeping in view the remarkable speed at which the
Fe2Al5 fingers grow into the steel (in HDA with pure aluminum), the lattice of Fe2Al5
must have a high concentration of defects, i.e., vacant atomic sites. And thus, any
increase in the lattice parameter of the Fe2Al5 phase can possibly be caused by an
increased occupation of the vacant aluminum-sites in the Fe2Al5 lattice. It appears that the
presence of silicon retards the diffusion through the interlayer and thus there is an
accompanying decrease in the rate at which the aluminum atoms are transferred to the
advancing interface, with the result that more aluminum tends to be retained in the lattice
of Fe2Al5 causing the lattice parameter to exhibit an increase.

Table 3.1. Variation in the lattice parameter of the Fe2Al5 phase with
the silicon content of the aluminum melt.
Silicon Level of the melt (%) d002 of the Fe2Al5 phase (A˚)
0.0 2.1158
1.8 2.1255
3.0 2.1500
6.0 2.1538

3.7 Bending of Aluminized Samples

Aluminized sheets of steels exhibit a very limited formability. The formability is


markedly affected by the thickness of interlayer. It is for this reason that where
formability is a requirement, the aluminizing is preferably carried out in silicon-
containing aluminum melts.
As discussed in the previous section, a gradually increasing silicon addition into the melt
causes a corresponding decrease in the thickness of the interlayer. In order to examine the
consequent effect on the formability, a test was designed which essentially comprised a
3-point bend test. The maximum angle of bend before cracking was taken as an index of
the ability of the aluminized strip to be formed or bent.

The bend test was conducted on an especially made device (or machine) for this purpose.
Photographs of the bending machine used during the present work are shown in Figure
3.20. For carrying out the test, a 2-inch long piece of strip cut from the aluminized
sample like the one shown in Figure 3.1 was subjected to the 3-point bending, in the
manner explained below:

1. The sample ‘S’ was placed in the bending machine in the specified space, as
shown in the photograph of Figure 3.20 (b), and the schematic illustration of
Figure 3.20 (c).
2. The Drive drum ‘D’ was rotated clock-wise so as to lift the loading-ram ‘F’ to a
level that it just touches the lower face of the sample.
3. The dial gauge ‘G’, and the graduated scale on the drive drum D, were then both
set to zero before starting the bending operation.
4. The bending was accomplished by rotating the drive D clockwise, while closely
observing the upper face of the sample with a magnifier of power x10, through
the window ‘W’ located on the top side of the machine.
5. The bending was continued until the stage that first cracks were observed to form
on the upper face of the sample.
(a)
Window (W)

Dial gauge (G)

Loading ram (F)

Drum (D)

Graduated
scale

Cont…..
(b)

Window (W)

Loading
ram (F)

Loading
ram (F)
Sample (S)

(c)

Figure 3.20. (a) Bending machine used for carrying out the bend type test on
aluminized samples, (b) the top view of the machine showing the ‘window
through which the cracking on the specimen surface could be monitored with a
magnifier, and (c) the schematic illustration of the 3-point bend test.
6. The displacement of the loading-ram F was then noted both from the dial gauge
G, as well as the drive drum D.
7. This displacement, which was related to the angle of the bend, was then taken as
an index of the formability or the ductility of the aluminized flat.

The results obtained from the bending of the samples aluminized in different melts were
plotted against the silicon-content of the aluminum melt. The graph shown in Figure 3.21
summarizes the behavior of the samples aluminized with different silicon contents of the
aluminum melts.

It can be clearly seen in Figure 3.21 that with an increase in the silicon-concentration of
the aluminum-melt, the angle of bend at which the start of the cracking was observed,
increased. It must be remembered that, as shown in the previous section, the additions of
silicon into aluminum cause a decrease in the thickness of the interlayer. Accordingly, it
would follow that the thinner the interlayer, the greater shall be the angle to which the
aluminized flats or sheets could be bent without cracking.
Figure 3.21. Graph showing relationship between silicon contents (wt%) and the
deflection in (mm).

(a)

(b)

cracks
Figure 3.22. (a) Showing the start of the cracking during the bend-test,
(b) cracking as observed with a magnifier.

The start of the cracking during the bend-test, as observed with a magnifier, is shown in
Figure 3.22. It may be noticed that the cracking essentially constituted the formation a
large number of cracks which were more or less parallel to each other, and were oriented
perpendicular to the direction of tensile stress established in the surface layer of the
specimen by the applied bending moment.

Sections were taken from the bent samples at right angle to the cracks so as to examine
the morphology and the distribution of cracks and their relationship with the interlayer.
The cracking pattern in the samples aluminized in pure aluminum was distinctly different
from that observed in the samples aluminized in Al-Si alloys.

The outside and the inside surface of the bent samples aluminized in pure aluminum are
shown in Figure 3.23 and Figure 3.24 respectively. It may be noted in Figure 3.23 that, at
the outside of the bend, the brittle Fe2Al5 layer has cracked through its entire depth, while
the cracks have not been able to penetrate into the tougher steel substrate. It was also
noted that at greater bend angles there were signs of deformation in the grains of the
substrate at the tips of the cracks. This observation clearly shows the presence of stress-
concentration at the crack-tips.

It was also observed that, in general, as the bend angle was gradually increased, the
opening of the already-existing cracks increased instead of the formation of new cracks.
This effect can be appreciated through a comparison between the micrographs given in
Figure 3.23 (a) and (b); in (a) the cracks are narrower but have penetrated the entire depth
of the interlayer; while in (b) the opening of the cracks has increased and yet no extension
of these cracks into the steel substrate can be evidenced. Nevertheless, it is possible that
at larger bend angle, new cracks may also start to form in addition to the pre-existing
cracks.

(a)

50 μm

(b)
50 μm

Cont…..

(c)

100 μm
Figure 3.23. Microstructures showing the outside surface of the bent samples
aluminized in pure aluminum.

The pattern of the cracking of brittle interlayer on the inside of the bent samples was
quite different from that on the outside of the bend. As shown in Figure 3.24 (a and b), on
the inside of the bend, the cracking had occurred through the interlayer and parallel to
sheet surface causing the aluminum coating to separate or ‘flake-off’ from the substrate.
Also, at greater bend angles, the aluminum coating tended to buckle outwardly after
separating from the substrate, Figure 3.24 (c and d).

The behavior of the aluminized sheets as observed through the metallographic


examination of the bent samples clearly showed that the interlayer is highly susceptible
to cracking whether it is stressed in tension (on the outside of the bend), or compression
(on the inside of the bend). It should also follow from the above observations that a
predictable and reliable formability cannot be expected from the aluminized sheets. It
shall be difficult to ensure that during the forming operation the interlayer does not
develop any cracking. However, it can also be debated that even if minor cracking had
occurred in the interlayer, as long as the outer aluminum coating is not damaged the
sheets (or the formed shape) may be acceptable for most applications.

It may however be appreciated that the cracking of the interlayer during forming should
also be affected by the thickness of the aluminized sheets, i.e., the thinner sheets are
expected to exhibited a relatively greater ability to be bent or formed before the interlayer
appears to start cracking.

Bending experiments conducted on flat samples aluminized in Al-6%Si alloy melts


showed interesting results. As may be seen in the metallographic section taken across the
cracks, given in Figure 3.25, that at many places the cracks had formed in the inter-
metallic compound layer while the outer aluminum layer is still quite ‘healthy’ and
continuous.

(a)

50 μm

Cont…..
(b)

25 μm

(c)

100 μm

Cont…..
(d)

50 μm

Figure 3.24 (a-d). Microstructures showing the inside surface of the bent
samples aluminized in pure aluminum.

It can be appreciated that during the bend test carried out on the bending device shown in
Figure 3.20, the cracking in the interlayer may have occurred much earlier than it was
indicated at the surface of the outer aluminum layer. Accordingly, the results of the bend
test as presented in Figure 3.21, may appear debatable. Nevertheless, these sub-surface
cracks can be the source of stress-concentration and thereby affect the overall formability
of the aluminized sheets.

It must be expected that the early cracking of the interlayer, and the stress-concentration
caused therefrom may also affect the tensile strength of the steel upon aluminizing. The
reduction in the tensile strength of steel wires and sheets upon aluminizing has been
understood [1,2,90] to be caused by the ‘tempering’ of the steel during the dipping into
molten aluminum above 650°C. However, on the basis of the present work, it appears
that the real cause of the deterioration of the tensile strength lies in the early cracking of
the interlayer.

(a)

25 μm

(b)

Cracks in the
interlayer
Figure 3.25 (a & b). Microstructures showing the outside surface of the bent
samples aluminized in Al-Si alloy. The cracks in the interlayer are indicated by
the arrowheads.

Further, as reported in the published literature [1,90] the wires and sheets coated in
aluminum-silicon alloys exhibit relatively better drawability and fabricability compared
to the wires and sheets aluminized in pure aluminum. The present work provides support
to these observation reported by ASM Committee [1]. The present work clearly shows
that smaller the thickness of the interlayer the smaller would be the length of the cracks in
the interlayer and therefore the smaller shall be the consequent degree of stress-
concentration. And thus as a result, the drawability and the fabricability shall also show
an improvement.

To sum up, it must be stated that the presence of the brittle layer of inter-metallic
compound just beneath the surface can affect, by means of early cracking and stress-
concentration, various mechanical characteristics including the strength, the drawability,
the fabricability, and possibly the weldability.

The behavior of the inside of the bent samples aluminized in Al-6%Si alloys is shown in
Figure 3.26. It may be noted that similar to the samples aluminized in pure aluminum
(Figure 3.24), the cracks are formed parallel to the sheet surface. The cracking occurs
through the interlayer causing the outer aluminum coat to separate or flake-off from the
substrate and buckle outwardly.

The ‘flaking’ of the aluminized sheets during forming or fabrication has also been
mentioned in the published literature [2,90]. It is logical to believe that such flaking must
have been observed at those locations where compressive stresses had developed during
the fabrication. The present work clearly shows that whereas the appearance of ‘cracks’
on the surface is related to the tensile stresses, the ‘flaking’ appears to be related to the
compressive stresses.

(a)

200 μm
(b)

cracked
interlayer

100 μm

Figure 3.26 (a & b). Microstructutres showing the inside surface of the bent
samples aluminized in Al-Si alloy. The locations where the aluminum coating
has buckled/flaked-off are indicated by the arrowheads. In Figure 3.26b, the
cracking through the intermetallic compound may also be noted.

3.8 Effects of ‘over-modification’ of Aluminizing Melts

The flux used for covering the aluminizing melts usually contains substantial quantities
of sodium. The flux used during the present work (see section 2.3 above) contained
chloride and fluoride of sodium.

In high silicon alloys of aluminum, this sodium can refine the distribution of silicon in the
aluminum matrix, an effect called ‘modification’ [91]. The ‘modified’ distribution of
silicon precipitates brings about an improvement in the strength as well as ductility of the
Al-Si alloys [91]. However, it has also been observed that when the concentration of
sodium (or strontium, another modifier) exceeds beyond a certain limit the ductility starts
to decrease again, and the alloy becomes ‘over-modified’, with a consequent loss in
ductility [91,92,93].
It has been observed that upon over-modification, isolated regions containing large plate-
like precipitates of silicon start to appear in a well modified fine eutectic matrix. These
plate-like precipitates can be of the order of 100 microns in size and exhibit a variety of
orientations with the aluminum matrix [92].

The bending experiments conducted during the present work, showed that the samples
aluminized in high-silicon melts showed a fairly inconsistent behavior in terms of the
‘angle of bend’ before the start of cracking. Accordingly, it was of interest to explore
whether the ‘over-modification’ of silicon precipitates inside the aluminum coating was
responsible for the inconsistency of results observed during the Bend-Test.

Microstructure given in Figure 3.27 shows the distribution of silicon precipitates in the
outer aluminum layer. It may be noted that whereas most of the silicon precipitates
exhibit a fine and uniform distribution, some precipitates are relatively quite large as
compared with the others. Further, some precipitates have formed as large plates and that
these plates exhibit random orientations. Also, that some of these large plates may be
relatively thicker compared with the rest.
(a)

primary Si
precipitates

25 μm

(b)

primary Si
precipitates

Figure 3.27 (a & b). Microstructures showing the randomly oriented primary
silicon precipitates inside the outer Al-Si alloy coat.
It was also noted that the large plates of silicon may sometimes be so oriented that during
the bend-test these plates would be perpendicular to the applied tensile stress. The
micrograph given in Figure 3.27 shows that some of the plates are oriented almost normal
to the sheet surface such that these plates can easily develop cracks under a tensile force
applied parallel to the sheet surface and perpendicular to the precipitate plates.

The microstructures of sections taken from the bent samples aluminized in ‘over-
modified’ melts are given in Figure 3.28. These microstructures have been purposely
taken from such locations where the cracking had occurred through the large silicon
plates. It may be noted that whereas the interlayer had cracked at many locations, the
cracks which had reached the outer surface, had propagated through the large silicon
precipitates. It can be easily appreciated from these observations that when large plates of
silicon are present in the outer aluminum-coat (as in the over-modified case) the cracks
shall show up at the surface earlier than when there were no large silicon precipitates
present.

(a)
cracked Si
precipitate

Cont…..
(b)

Figure 3.28 (a & b). Scanning electron micrographs of bent samples showing
the preferential propagation of cracks through the primary Si precipitates, as
indicated by the arrowheads.

In order to investigate in detail the behavior of the ‘large’ silicon plates in the over-
modified Al-Si alloys, samples were taken from the aluminizing melts and cast into bars
which were then tested in tensile mode. Selected broken sample were then sectioned
longitudinally to examine the path of the fracture crack.

Fig. 3.29 shows the microstructures of Al-13%Si alloy in the modified and over-modified
conditions. It may be noted that in the properly modified alloy (Figure 3.29 a), the
precipitates exhibit a fine and uniform distribution in the eutectic regions of the
microstructures. The white areas in these microstructures are the pro-eutectic aluminum
phase. It may also be noted that in the over-modified microstructure (3.29 b), in addition
to the colonies of finely-distributed silicon precipitates, a number of large plate-like
precipitates of silicon have also formed. These plate-like silicon precipitates are typically
about 100 microns in size and exhibit different orientations in space. Also, some of the
plates are relatively ‘thicker’ than most of the other large plate-like precipitates.

(a)

50 μm

(b)

50 μm
Figure 3.29 Microstructures of the Al-13%Si alloys in the (a) modified,
(b) over-modified conditions. Large plate-like precipitates of primary silicon
may be noticed in the over-modified microstructure of Figure 3.29 b.

Test-bars made from modified and over-modified melts were broken in the tensile mode
and were examined microscopically. Scanning electron micrographs of the fracture face
[92,94] of modified and over-modified samples are given in Figure 3.30. Typical
character of a ‘ductile’ (Figure 3.30a), and a ‘brittle’ (Figure 3.30b) fracture can be
evidenced in these micrographs.

(a)
(b)

Figure 3.30 (a) Scanning electron micrograph of a modified alloy showing ductile
fracture, (b) Scanning electron micrograph of an over-modified alloy showing brittle
cleavage fracture [92,94].

Test-bars broken in tensile mode were also sectioned longitudinally and examined under
an optical microscope. Selected micrographs taken from the over-modified alloy are
given in Figure 3.31 It is important to point out that all the microstructures of the broken
samples presented here have been placed on the page such that the tensile axis was
oriented vertically in the plane of the page.

In Figures 3.31 (a) and (b) the crack can be seen to have traveled through a plate-like
precipitate of silicon. It can also be seen that the ‘cracked’ plate-like precipitates are
oriented perpendicular to the tensile axis, while those plate-like precipitates which are not
perpendicular to the tensile axis did not develop any cracking. In Figures 3.31 (c) and (d)
the cracking appears to be confined to the thicker precipitates, although the cracking
within the precipitates has still occurred normal to the tensile force.

An important feature that must be carefully noted is that it is the precipitate plates which
have cracked, and not the sharp corners of the precipitates which might have been
expected to act as the crack origin. This observation is in contradiction to the view that
exists in the published literature [95] according to which the sharp corners or edges of the
primary silicon precipitates act as crack nuclei. It is also clearly evident that the thicker
precipitates were easier to crack.

Microstructures given in 3.31 (a) and (c) have been taken from a region near the surface
of the specimen. These microstructures clearly show that the crack had passed through
the body of a precipitates which were oriented perpendicular to the tensile axis. It is once
again clear from these micrographs that it were not the sharp edges of the precipitate
which were responsible for the initiation of the crack through some ‘stress concentration’.

The above analysis clearly indicates that the behavior expected from large and/or thick
primary silicon precipitates in the aluminum coating layer would be to assist the
propagation of the cracks. The presence of such precipitates, therefore, can reduce the
affective bend-angle at which the cracking would start to appear at the surface of the
specimens during the bend test.
(a)

100 μm

(b)

50 μm

Cont…..
(c)

50 μm

(d)

50 μm
Figure 3.31. Microstructures taken from the longitudinal section of a broken tensile
specimen showing the preferential propagation of the cracks through the primary
silicon plates oriented perpendicular to the tensile axis. The double sided arrowheads
indicate the direction of the applied tensile force. It may also be noted in Figure
3.31(d) that the ‘thicker’ precipitates appear more susceptible to cracking.

4. Conclusions

In aluminizing with pure aluminum, the interlayer that forms between the aluminum and
the substrate steel consists of orthorhombic Fe2Al5 phase. The interlayer grows into the
steel substrate as finger-like columnar crystals, oriented perpendicular to the surface
being aluminized. These columnar finger-like crystals exhibit a preferred lattice
orientation such that the [002]orth is normal to the substrate surface and along the
longitudinal axis of the columnar crystals.

Also, as the interlayer advances into the steel, the carbon in the steel diffuses ahead of the
advancing interface causing a ‘carbon build-up’, with the result that an increased volume
fraction of pearlite is observed next to the interlayer.

With the additions of silicon into the molten aluminum, the finger-like character of the
interlayer starts to disappear resulting in a comparatively smoother boundary between the
interlayer and the substrate. The interface becomes almost ‘planar’ at a silicon level of
about 6 %. The carbon build-up is also observed in these samples, however, the pearlite
phase forms as a narrow and continuous band along the interface.

XRD and SEM studies show that when silicon is present, the interlayer actually consists
of two layers. The Fe2Al5 layer next to the substrate (with a finger-like morphology at
low-silicon-levels) gradually reduces in thickness as the silicon-level is increased, and the
outer Fe3Al layer (adjacent to the outer aluminum-coat) gradually builds up in thickness.

Bending experiments conducted on aluminized samples showed that the presence of the
brittle layer of inter-metallic compound just beneath the surface can affect, by means of
early cracking and stress-concentration, various mechanical characteristics including the
strength, the drawability, the fabricability, and possibly the weldability.

5. Suggestions for future work

1. Efforts may be made either to extract the interlayer from the samples aluminized
in aluminum-silicon alloy (with 6% or higher silicon), or to obtain a powder
sample of the interlayer as was accomplished during the present work. Such an
exercise shall serve to confirm the presence of Fe3Al phase or otherwise. It must
be realized that in the aluminized steel samples, the higher intensity peaks of
Fe3Al are overlapped by either the aluminum peaks or the iron peaks, which
makes it difficult to confirm with certainty the presence of Fe3Al. It must also be
realized that any identification based on electron probe micro-analysis can not be
as conclusive (due to the interference of the neighboring phases) as the X-ray
diffraction.
2. During the present work no evidence for the presence of FeAl3 phase in the
interlayer was observed. The previous works have also not presented any
convincing evidence. More work needs to be carried out to confirm the presence
or absence of this phase.
3. The modification of the morphology of interface between the coating and the
substrate may also be tried through the additions of the elements other than
silicon. Such an exercise may prove quite useful considering that when silicon is
used the primary crystals of silicon (in over-modified alloys) deteriorate the
ductility of the coating.
Appendix - I
_______________________________________________________________

Name and formula

Reference code: 00-014-0336

PDF index name: Aluminum Iron

Empirical formula: Al5Fe2

Chemical formula: Al5Fe2

Crystallographic parameters

Crystal system: Orthorhombic


Space group: Cmcm
Space group number: 63

a (Å): 7.6750
b (Å): 6.4030
c (Å): 4.2030
Alpha (°): 90.0000
Beta (°): 90.0000
Gamma (°): 90.0000

Volume of cell (10^6 pm^3): 206.55


Z: 2.00
RIR: -

Status, subfiles and quality

Status: Marked as deleted by ICDD


Subfiles: Inorganic
Alloy, metal or intermetalic
Quality: Blank (B)

Comments

Deleted by: Deleted by 29-43.


General comments: Single-crystal data taken.
Actual composition is Al5.14Fe2 .

References

Primary reference: Hanawalt et al., Anal. Chem., 10, 475, (1938)


Unit cell: Schubert et al., Naturwissenschaften, 40, 437, (1953)

Peak list

No. h k l d [A] 2Theta[deg] I [%]


1 1 1 0 4.90000 18.089 12.0
2 2 0 0 3.86000 23.022 25.0
3 0 2 0 3.20000 27.858 40.0
4 3 1 0 2.39000 37.604 10.0
5 0 0 2 2.11000 42.824 100.0
6 1 3 0 2.05000 44.142 100.0
7 1 1 2 1.94000 46.789 10.0
8 4 0 0 1.90000 47.835 8.0
9 2 0 2 1.84000 49.498 4.0
10 0 2 2 1.76000 51.911 8.0
11 1.70000 53.888 2.0
12 3 3 0 1.63000 56.403 2.0
13 2 2 2 1.59000 57.955 4.0
14 4 2 1 1.55000 59.599 2.0
No. h k l d [A] 2Theta[deg] I [%]
15 3 3 1 1.52000 60.899 10.0
16 2 4 0 1.48000 62.728 16.0
17 4 0 2 1.42000 65.703 2.0
18 2 4 1 1.39000 67.307 10.0
19 1 1 3 1.35000 69.583 2.0
20 4 2 2 1.30000 72.675 2.0
21 0 4 2 1.27000 74.679 10.0
22 5 3 0 1.24000 76.809 8.0
23 1 5 1 1.21000 79.079 16.0
24 4 4 1 1.18000 81.506 2.0
25 3 5 0 1.15000 84.107 2.0
26 3 5 1 1.10000 88.898 8.0
27 6 0 2 1.09000 89.934 2.0
28 5 3 2 1.07000 92.094 10.0
29 2 6 0 1.03000 96.811 4.0
30 5 1 3 1.02000 98.085 2.0

Stick Pattern
Appendix - II
_______________________________________________________________

Name and formula

Reference code: 00-045-1205

PDF index name: Aluminum Iron Silicon

Empirical formula: Al0.5Fe3Si0.5

Chemical formula: Al0.5Fe3Si0.5

Crystallographic parameters
Crystal system: Cubic
Space group: Fm-3m
Space group number: 225

a (Å): 5.7210
b (Å): 5.7210
c (Å): 5.7210
Alpha (°): 90.0000
Beta (°): 90.0000
Gamma (°): 90.0000
Calculated density (g/cm^3): 6.92

Volume of cell (10^6 pm^3): 187.25


Z: 4.00
RIR: -

Subfiles and Quality

Subfiles: Inorganic
Alloy, metal or intermetalic
Quality: Calculated (C)

Comments

Unit cell: The lattice parameter for Al0.6Fe3Si0.4 : a=5.7357;


Al0.4Fe3Si0.6 : a=5.7074.

References
Primary reference: Hubbard, C., Oak Ridge National Laboratory, High Temp.
Mat. Lab., Oak Ridge, TN, USA., Private
Communication, (1993)
Unit cell: Cowdery, S., Kayser, F., Mater. Res. Bull., 14, 91, (1979)

Peak list

No. h k l d [A] 2Theta[deg] I [%]


1 1 1 1 3.30300 26.973 7.0
2 2 0 0 2.86100 31.238 3.0
3 2 2 0 2.02300 44.763 100.0
4 3 1 1 1.72490 53.049 2.0
5 2 2 2 1.65150 55.605 1.0
6 4 0 0 1.43030 65.171 12.0
7 3 3 1 1.31250 71.874 1.0
8 4 2 0 1.27930 74.045 1.0
9 4 2 2 1.16780 82.541 19.0
10 4 4 0 1.01130 99.227 6.0
11 6 2 0 0.90460 116.758 8.0
12 4 4 4 0.82580 137.749 2.0

Stick Pattern
7. References
[1] ASM Handbook committee; Aluminum coating of Steel, ASM metals Hand Book, on
Surface Cleaning, Finishing and Coating, ASM International, Material Park, OH,
5(1982) 335-347.

[2] AK Steel; Technical Bulletin on Aluminized Steel Type 2, AK Steel Corporation, 703,
Curtis Street, Middle Town, OH.

[3] U.R. Kattner and B.P. Burton; Binary Alloy Phase Diagrams, ASM Metals
Handbook, on Alloy Phase Diagram, ASM International, Material Park, OH, 3(1999)
2-44.

[4] A. G. C. Gwyer; Z. anorg. Chem., 57(1908) 126-133.

[5] N. S. Kurnakow, G. Urasow and A. Grigorjew; Z. anorg. Chem., 125(1922) 207-227.

[6] M. Isawa and T. Murakami, Kinzoku-no-Kenkyu, 4(1927) 467-477.

[7] A. G. C. Gwyer and H. W. L. Phillips; J. Inst. Metals, 38 (1927) 35-44 & 83.

[8] F. Wever and A. Miiller; Z. anorg. Chem., 192(1930) 340-345.

[9] W. Oelsen; Stahl u. Eisen, 69(1949) 468-474.

[10] N. W. Ageew and O. I. Vher; J. Inst. Metals, 44(1930) 84-85.

[11] W. Ekman; Z. physik. Chem., 1312(1931) 57-78.

[12] A. J. Bradley and A. H. Jay; Proc. Roy. Soc. (London), A136(1932) 210-232.

[13] A. J. Bradley and A. H. Jay; J. Iron Steel Inst., 125(1932) 339-357.

[14] C. Sykes and H. Evans; Proc. Roy. Soc. (London), A145(1934) 529-539.

[15] C. Sykes and H. Evans; J. Iron Steel Inst., 131(1935) 225-247.

[16] A. T. Grigoriev and N. M. Gruzdeva; Izvest. Sektora Fiz.-Khim. Anal., Inst.


Obshchei Neorg. Akad. Nauk S.S.S.R., 14(1941) 245-253.
[17] H. Saito; Nippon Kinzoku Gakkai-Shi, 14/5(1950) 1-6 & 6-11.

[18] W. Sucksmith; Proc. Roy. Sac. (London), A171(1939) 525-540.

[19] W. D. Bennett; J. Iron Steel Inst., 171(1952) 372-380.


[20] M. Fallot; Ann. phys., 6(1936) 356-365.

[21] A. J. Bradley and A. Taylor; Proc. Roy. Soc. (London), A166(1938) 353-375.

[22] W. L. Bragg; Nature, 131(1933) 751.

[23] K. Schafer; Naturwissenschaften, 21(1933) 207.

[24] C. Sykes and J. W. Bampfylde; J. Iron Steel Inst., 130(1934) 389-410.

[25] W. H. Rothery and H. M. Powell; Z. Krist., 91(1935) 23-47.

[26] S. Matsuda; J. Phys. Soc. Japan, 6(1951) 131-135.

[27] H. Sato; Science Repts. Research Insts. Tohoku Univ., A3(1951) 13-23.

[28] M. Hansen; Constitution of Binary Alloys, McGraw-Hill Book Company, Second


Edition (1958) 108-115.

[29] A. Osawa; Science Repts. Tohoku Univ., 22(1933) 803-819.

[30] A. J. Bradley and A. Taylor; J. Inst. Metals, 66(1940) 53-65.

[31] A. G. C. Gwyer, H. W. L. Phillips and L. Mann; J. Inst. Metals, 40(1928) 302 &
358.

[32] P. J. Black; Acta Cryst., 8(1955) 175-182.

[33] G. V. Raynor, C. R. Faulkner, J. D. Noden and A. R. Harding; Acta Met., 1(1953)


629-635.

[34] M. Armand; Compt. rend., 235(1952) 1506-1508.

[35] G. V. Raynor and P. C. L. Pfeil; J. Inst. Metals, 73(1946-1947) 397-419.

[36] H. Nowotny, K. Komarek and J. Kromer; Berg. u. Huttenmdnn. Monatsh., 96(1951)


161-169.

[37] E. Gebhardt and W. Obrowski; Z. Metallkunde, 44(1953) 154.

[38] R. S. Archer and W. L. Fink; J: Inst. Metals, 40(1928) 356-357.

[39] W. Rosenhain, S. L. Archbutt and D. Hanson; Z. Metallkunde, 18(1926) 65.

[40] E. Wetzel; Metallborse, 13(1923) 738.


[41] E. H. Dix; Proc. Am. Soc. Testing Materials, 25/2(1925) 120-129.

[42]H. Hanemann and A. Schrader; Z. Metallkunde, 31(1939) 183.

[43] G. Masing and O. Dahl; Z. anorg. Chem., 154(1926) 189-196.

[44] A. Roth; Z. Metallkunde, 31(1939) 299-301.


[45] J. K. Edgar; Trans. AIMS, 180(1949) 225-229.

[46] G. Falkenhagen and W. Hofmann; Z. Metallkunde, 43(1952) 69-81.

[47] Z. Nishiyama; Science Repts. Tohoku Univ., 18(1929) 381-387.

[48] A. Osawa; Kinzoku-no-Kenkyu, 10(1933) 432-445.

[49] K. Schubert and M. Kluge; Z. Naturforsch., 8a(1953) 755-776.

[50] E. Bachmetew; Z. Krist., 89(1934), 575-586; & 88(1934) 179-181.

[51] G. Phragmn; J. Inst. Metals, 77(1950) 489.

[52] P. J. Black; Acta Cryst., 8(1955) 43-48 & 175-182.

[53] H. J. Beattie and F. L. VerSnyder; Trans. ASM, 45(1953) 397-423.

[54] R.W. Richards, R.D. Jones, P.D. Clements and H. Clarke; Int. Mater, Rev. 39 (1994)
191–212.

[55] N.A. El-Mahallawy, M.A. Taha, M.A. Shady, A.R. El-Sissi, A.N. Attia and W. Rief;
Mater. Sci. Technol, 13(1997) 832–840.

[56] G. Eggeler, W. Auer and H. Kaesche; J. Mater. Sci., 21(1986) 3348–3350.

[57] M. Yu, R. Shivpuri and R.A. Rapp; J. Mater. Eng. Perform, 4 (1995) 175–181.

[58] V.I. Dybkov; J. Mater. Sci., 25(1990) 3615–3633.

[59] D.O. Gittings, D.H. Rowland and J.O. Mack; Trans. ASM, 43(1950) 587–610.

[60] C.M. Cotell, J.A. Sprague and F.A. Smidt; Surface Engineering, ASM Metals
Handbook, on Surface Engineering, ASM International, Materials Park, OH, 5(1999)
718.

[61] W. Deqing, S. Ziyuan and Z. Longjiang; Applied Surface Science, 214(2003) 304–
311.
[62] S. Kobayashi and Takao Yakou; Mater. Sci. Eng., A338(2002) 44-53.

[63] S. H. Hwang, J. H. Song and Y. S. Kima; Mater. Sci. Eng., A390(2005) 437–443.

[64] L. Yajiang, W. Juan, Z. Yonglan and X. Holly; Bull. Mater. Sci., 25/No. 7(2002)
635–639.

[65] C. J. Wang and S. M. Chen; Surface & Coatings Technology, 200(2006) 6601–6605.

[66] H.R. Shahverdi, M.R. Ghomashchi, S. Shabestari and J. Hejazi; J. of Materials


Processing Technology, 124(2002) 345-352.

[67] D.M. Dovey and A. Waluski; Metallurgia, (1963) 211–217.

[68] M.A. Shady, A.R. El-Sissi, A.M. Attia, N.A. El-Mahallawy, M.A. Taha and W.
Reif; J. Mater. Sci. Lett, 15(1996) 1032–1036.

[69] T. Morinaga and Y. Kato; J. Jpn. Inst. Metals, 19(1955) 578.

[70] S. Koda, S. Morozumi and A. Kanai; J. Jpn. Inst. Metals, 26(1962) 764.

[71] M.V. Akdeniz, A.O. Mekhrabov and T. Yilmaz; Scri. Metall. Mater., 31(1994)
1723.

[72] V.G. Gurtler and K. Sagel; Z. Metallkde, 46(1955) 738.

[73] T. Heumann and S. Dittrich; Z. Metallkde, 50(1959) 617.

[74] G. Eggeler, W. Auer and H. Kaesche; Z. Metallkde, 77(1986) 239.

[75] K. Bouche, F. Barbier and A. Coulet; Mater. Sci. Eng., A249(1998) 167.

[76] V. Stefaniay, A. Griger and T. Turmezey; J. Mater. Sci., 22(1987) 539–546.

[77] S.C. Kwon and J.Y. Lee; Can. Metall. Quart., 20(1981) 351–357.

[78] A. Knauscher; Neue Hiitte, 19(1974) 398.

[79] S. G. Denner, R. D. Jones and R. J. Thomas; Iron Steel Int., 48(1975) 241.
[80] J. E. Nicholls; Corr. Technol., 11(1964) 16.

[81] D. I. Lainer and A. K. Kurakin; Fiz. Met. Metallov., 18(1964) 145.

[82] N. Komatsu, M. Nakamura and H. Fujita; J. Jpn. Inst. Met., 45(1981) 416.
[83] R. D. Jones and S. G. Denner; Proceedings 10th World Congress in Metal Finishing,
(Metal Finishing Society of Japan, Kyoto) 1980.

[84] V. G. Rivlin and G. V. Raynor; Int. Met., Rev. 26(1981) 133.

[85] M.V. Akdeniz and A.O. Mekhrabov; Scri. Metall. Mater., 46(1998) 1185.

[86] H. Glasbrenner and E. Nold, Z. Voss; Journal of Nuclear Materials, 249(1997) 39-
45

[87] V.N. Yeremenko, Y.N. Natanzon and V.I. Dybkov; J. Mater. Sc., 16(1981) 1748.

[88] ICDD X-Ray File, Reference Code # 00-045-1204.

[89] F. Hasan, A. Jahanafrooz, G.W. Lorimer and N. Ridley; Met. Trans., A 13A(1982)
1337.

[90] AK Steel; Production Data Bulletin on Aluminized Steel Type 1, AK Steel


Corporation, 703, Curtis Street, Middle Town, OH.

[91] L. Backrud, G. Ghai and J. Tamminen; Solidification Characteristics of Aluminum


alloys, Foundry Alloys, AFS/SKANALUMINUM, 2(1990).

[92] G. H. Awan and F. Hasan; Pakistan Institute of Physics International Conference


Lahore, 2006.

[93] J. Gruzleski and B. Closset; The treatment of Aluminum Alloys, AFS (1990).

[94] K. B. Ahmad; M.Sc. Thesis, Modification of Microstructure and Mechanical


Properties of 413 Al-Si Eutectic Alloy, University of Engineering & Technology,
Lahore, 2004.

[95] A. Gangullee and J Gurland; Trans. Metall. Soc. AIME, 239/2(1967) 269-272

Potrebbero piacerti anche